University of Calgary PRISM: University of Calgary's Digital Repository

Graduate Studies The Vault: Electronic Theses and Dissertations

2016 Molecular mechanisms of anti-inflammatory glucocorticoid action: Evidence for a role of glucocorticoid-inducible and implications for glucocorticoid insensitivity

Shah, Suharsh Vinaykumar

Shah, S. V. (2016). Molecular mechanisms of anti-inflammatory glucocorticoid action: Evidence for a role of glucocorticoid-inducible genes and implications for glucocorticoid insensitivity (Unpublished doctoral thesis). University of Calgary, Calgary, AB. doi:10.11575/PRISM/26991 http://hdl.handle.net/11023/3002 doctoral thesis

University of Calgary graduate students retain copyright ownership and moral rights for their thesis. You may use this material in any way that is permitted by the Copyright Act or through licensing that has been assigned to the document. For uses that are not allowable under copyright legislation or licensing, you are required to seek permission. Downloaded from PRISM: https://prism.ucalgary.ca UNIVERSITY OF CALGARY

Molecular mechanisms of anti-inflammatory glucocorticoid action: Evidence for a role of

glucocorticoid-inducible genes and implications for glucocorticoid insensitivity

by

Suharsh Vinaykumar Shah

A THESIS

SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE

DEGREE OF DOCTOR OF PHILOSOPHY

GRADUATE PROGRAM IN CARDIOVASCULAR AND RESPIRATORY SCIENCES

CALGARY, ALBERTA

MAY, 2016

© Suharsh Vinaykumar Shah 2016 Abstract

The anti-inflammatory activity of glucocorticoids, in diseases such as asthma, is attributed to their ability to reduce the expression of multiple inflammatory mediators. While glucocorticoid

(NR3C1)-mediated transcriptional activation, or transactivation, was believed to primarily mediate side-effects, accumulating evidence indicates that transactivation is also important for the inhibition of inflammatory expression. This illustrates the need to investigate possible functional roles for specific glucocorticoid-inducible genes.

In this thesis, a repressive role for DUSP1, a phosphatase that inhibits MAPKs, was investigated.

DUSP1 was induced by IL1B and dexamethasone in both human pulmonary A549 and primary

HBE cells. IL1B-induced DUSP1 negatively regulated MAPK activity and in the further presence of dexamethasone, DUSP1 played a transient, typically partial, role in repressing expression of inflammatory genes, including CXCL1, CXCL2 and PTGS2. Thus, additional glucocorticoid- induced gene products are necessary for repression.

Regulation of the mRNA destabilizing , ZFP36, by DUSP1 was examined. Following the loss of DUSP1, ZFP36 expression was enhanced and this attenuated IL1B-induced TNF expression. Despite a modest ability of dexamethasone to induce ZFP36, thereby off-setting loss of ZFP36 due to reduced MAPK activity, neither silencing of dexamethasone-induced ZFP36,

DUSP1, nor both together, prevented repression of TNF by dexamethasone.

Finally, while DUSP1 over-expression attenuated IL1B-induced expression of many inflammatory genes (e.g. IL8, CSF2 etc.), others, including, the inflammatory , IRF1, and downstream target genes, such as CXCL10, were profoundly enhanced. While MAPK inhibition prolonged IRF1 expression, silencing of IL1B plus dexamethasone-induced DUSP1 reduced IRF1

ii and CXCL10 expression. Since, CXCL10 expression was largely unaffected by dexamethasone, these data suggest a mechanism whereby dexamethasone-induced DUSP1 expression maintains

CXCL10 expression.

In conclusion, this study demonstrates interlinked counterregulatory networks, in which IL1B- induced DUSP1 and ZFP36 regulate MAPK activation and inflammatory respectively. The transient and partial repressive effect of dexamethasone-induced DUSP1 on only few IL1B-induced inflammatory genes provides a rational for functional screening of glucocorticoid-inducible genes that may show repressive functions. Furthermore, by switching off

MAPKs, DUSP1 maintains IRF1 expression and may contribute to glucocorticoid insensitivity.

iii Acknowledgements

I would like to take this opportunity to thank many people who have travelled with me along this journey of PhD and making it a challenging but a zestful experience right from the very beginning. This thesis wouldn’t have been possible without the guidance and support of my supervisor, Dr. Robert Newton, whose endless inputs and advice throughout the process has led me to where I am today. I must say that when times were hard and the results were not that pleasing, he pushed my limits to success in times of failure. He undoubtfully had faith in the new ideas that I brought to the table and always welcomed my opinions and suggestions.

I would like to also express my great regards to the whole team of the Newton and Giembycz lab, especially to Liz, Neil, and Chris for creating a wonderful training and learning environment for moulding my scientific career. I also express my sincere thanks Mahmoud and Omar for all their help and support during the last few months of my PhD. Also, extra thanks to the members of the Proud and Leigh lab, especially Shahina, Suzanne and Cora for their endless support throughout the course of my PhD.

A very big thanks to the members of my committee, Dr. Proud and Dr. MacNaughton for their constant feedbacks and comments in shaping my career in the right direction. It has been a remarkable experience working alongside such esteemed pioneers of research and their invaluable commitments towards my research has been instrumental.

I would also like to also express my sincere gratitude towards my funding agencies Alberta Lung Association, Alberta Innovates – Health Solutions, the University of Calgary, Allergen etc. namely for the funding they have provided me to nurture my PhD project and scientific career.

Finally, I would like to thank my wife, Lino, my parents, family and friends for their undeviating support during this important phase of my career. This wouldn’t have been possible without you all.

iv Dedication

I would like to dedicate this thesis to my parents and grandparents. I believe that this achievement wouldn’t have been possible without their sincere devotion towards providing me with the best quality education.

v Table of Contents

Abstract ...... ii Acknowledgements ...... iv Dedication ...... v Table of Contents ...... vi List of Tables ...... x List of Figures ...... xi List of Abbreviations ...... xiv

CHAPTER 1: INTRODUCTION ...... 1 1.1 Asthma ...... 1 1.1.1 Burden of asthma ...... 2 1.1.2 Asthma pathogenesis ...... 3 1.1.3 Exacerbation of asthma ...... 8 1.1.4 Pulmonary epithelium and its role in asthma ...... 9 1.1.5 Current asthma therapy ...... 11 1.2 Control of inflammatory gene expression ...... 14 1.2.1 Mitogen activated protein kinase (MAPK) pathways ...... 14 1.2.1.1 Role of MAPK pathways in inflammation and asthma ...... 17 1.2.2 Transcription ...... 19 1.2.3 Inflammatory transcription factors ...... 21 1.2.3.1 NF-B ...... 22 1.2.3.2 IRF1 ...... 25 1.2.4 Post-transcriptional control of gene expression...... 27 1.2.4.1 mRNA stability ...... 27 1.3 Glucocorticoids ...... 29 1.3.1 Endogenous glucocorticoids ...... 29 1.3.2 Molecular mechanisms of glucocorticoid action ...... 30 1.3.3 Transrepression by glucocorticoids ...... 32 1.3.4 Transactivation by glucocorticoids ...... 34 1.3.4.1 DUSP1 ...... 35 1.3.4.2 ZFP36 ...... 37 1.3.5 Insensitivity/resistance to glucocorticoids in asthma ...... 39 1.3.5.1 Mechanisms of reduced glucocorticoid sensitivity ...... 39 1.4 General hypothesis and research aims ...... 42 1.4.1 General hypothesis ...... 42 1.4.2 Research aims: ...... 42

CHAPTER 2 : MATERIAL AND METHODS...... 43 2.1 Materials ...... 43 2.2. Methods ...... 45 2.2.1 Cell Culture methods ...... 45 2.2.1.1 A549 cell culture ...... 45 2.2.1.2 Primary human bronchial epithelial (HBE) cell culture ...... 45

vi 2.2.1.3 Adenoviral infection ...... 46 2.2.1.4 siRNA-mediated gene silencing ...... 46 2.2.2 Western blotting ...... 47 2.2.2.1 Preparation of cell lysates ...... 47 2.2.2.2 SDS polyacrylamide gel electrophoresis ...... 47 2.2.2.3 Protein transfer to nitrocellulose membranes ...... 47 2.2.2.4 Immunodetection of ...... 48 2.2.3 RNA isolation, cDNA synthesis and SYBR Green Real Time PCR ...... 48 2.2.3.1 RNA isolation ...... 48 2.2.3.2 cDNA synthesis ...... 49 2.2.3.3 SYBR green real-time PCR ...... 50 2.2.3.4 Analysis of unspliced nuclear RNA ...... 50 2.2.4 Enzyme-linked immunosorbent assay (ELISA) ...... 51 2.2.5 Chromatin Immunoprecipitation (ChIP) assay ...... 52 2.2.5.1 Cell fixation ...... 52 2.2.5.2 Chromatin shearing by sonication ...... 53 2.2.5.3 Immunoprecipitation of sheared chromatin ...... 54 2.2.5.4 Washing and cross-link reversal of ChIP DNA ...... 55 2.2.5.5 ChIP DNA clean-up ...... 55 2.2.5.6 SYBR green PCR using ChIP DNA ...... 56 2.2.6 3-(4, 5-dimethylthiazol-2-yl)-2, 5-diphenyltetrazolium bromide (MTT) assay ...... 56 2.2.7 Data presentation and statistical analyses ...... 57

CHAPTER 3 : FEEDBACK CONTROL OF INFLAMMATORY GENE EXPRESSION BY THE MITOGEN-ACTIVATED PROTEIN KINASE (MAPK) PHOSPHATASE, DUSP1, AND REGULATION BY DEXAMETHASONE ...... 58 3.1 Rationale ...... 59 3.2 Hypothesis ...... 60 3.3 Results ...... 60 3.3.1 Effect of dexamethasone on IL1B-induced MAPK activation and DUSP1 expression ...... 60 3.3.2 Effect of dexamethasone on IL1B-induced gene expression ...... 62 3.3.3 Effect of MAPK inhibitors on IL1B-induced gene expression ...... 64 3.3.4 Effect of DUSP1 over-expression on MAPK activation and inflammatory gene expression ...... 70 3.3.5 Effect of DUSP1 siRNA on MAPK activation by IL1B in the absence and presence of dexamethasone ...... 72 3.3.6 Effect of DUSP1 knock-down on IL1B-induced inflammatory gene mRNA expression in the absence and presence of dexamethasone...... 76 3.3.7 Effect of DUSP1 knock-down on IL1B-induced protein expression in the absence and presence of dexamethasone...... 82 3.4 Discussion ...... 85

CHAPTER 4 : DOWNREGULATION OF THE MITOGEN-ACTIVATED PROTEIN KINASE PHOSPHATASE, DUSP1 LIMITS TNF EXPRESSION THROUGH ENHANCED EXPRESSION OF TRISTETRAPROLIN (TTP; ZFP36) ...... 91

vii 4.1 Rationale ...... 92 4.2 Hypothesis ...... 93 4.3 Results ...... 93 4.3.1 Characterization of TNF expression in the presence of IL1B and dexamethasone.... 93 4.3.2 Analysis of DUSP1 expression in the presence of IL1B and dexamethasone ...... 97 4.3.3 Analysis of ZFP36 expression in the presence of IL1B and dexamethasone ...... 98 4.3.4 Effect of MAPK inhibitors on DUSP1 and ZFP36 expression ...... 98 4.3.5 Analysis of DUSP1 expression in the presence of IL1B and dexamethasone in primary HBE cells ...... 100 4.3.6 Analysis of ZFP36 expression in the presence of IL1B and dexamethasone in primary HBE cells ...... 101 4.3.7 Characterization of TNF expression in the presence of IL1B and dexamethasone in HBE cells ...... 102 4.3.8 Effect of ZFP36 siRNA on IL1B-induced TNF mRNA expression ...... 104 4.3.9 Effect of DUSP1 over-expression on ZFP36 and TNF mRNA expression ...... 107 4.3.10 Effect of MAPK inhibitors on TNF expression...... 108 4.3.11 Effect of DUSP1 knock-down on ZFP36 protein and TNF mRNA expression ..... 111 4.3.12 Role of ZFP36 in the loss of TNF mRNA following DUSP1 knock-down ...... 113 4.3.13 Effect of MAPK inhibitors and DUSP1 over-expression in the regulation of TNF protein expression ...... 116 4.3.14 Effect of DUSP1- and ZFP36-targeting siRNAs in the regulation of TNF protein expression ...... 117 4.4 Discussion ...... 119

CHAPTER 5: MAINTENANCE OF IRF1 BY DUSP1 ENHANCES IRF1-DEPENDENT GENE EXPRESSION: IMPLICATIONS FOR GLUCOCORTICOID THERAPY ...... 128 5.1 Rationale ...... 128 5.2 Hypothesis ...... 129 5.3 Results ...... 129 5.3.1 Effect of DUSP1 over-expression on inflammatory gene expression ...... 129 5.3.2 Kinetics of IL1B-induced inflammatory mRNAs ...... 131 5.3.3 Identification of IRF1-dependent mRNAs induced by IL1B ...... 132 5.3.4 Effect of DUSP1 over-expression on IL1B-induced IRF1 expression...... 133 5.3.5 Effect of MAPK inhibitors on IRF1 expression ...... 134 5.3.6 Effect of MAPK inhibitors on IRF1 transcription and mRNA stability...... 137 5.3.7 Effect of MAPK inhibitors on IRF1 protein stability ...... 138 5.3.8 Characterization of IRF1 expression in the presence of IL1B and dexamethasone . 140 5.3.9 Effect of dexamethasone on IRF1 transcription rate ...... 142 5.3.10 Effect of dexamethasone on IRF1 mRNA and protein stability ...... 144 5.3.11 Effect of dexamethasone on IRF1-dependent gene expression ...... 144 5.3.12 Effect of DUSP1 silencing on MAPKs and IRF1 expression ...... 145 5.3.13 Effect of DUSP1 silencing and MAPK inhibitors on IRF1-dependent gene expression ...... 147 5.3.14 Role of IRF1 in late-phase gene expression in the presence of IL1B and IL1B plus dexamethasone ...... 148

viii 5.3.15 Occupancy of IRF1 at the promoters of IRF1-dependent genes in the presence of IL1B and IL1B plus dexamethasone ...... 151 5.4 Discussion ...... 152

CHAPTER 6 : DISCUSSION ...... 160 6.1 Feedback and feed-forward control of inflammatory gene expression ...... 160 6.2 DUSP1-mediated regulation of inflammatory gene expression ...... 163 6.3 Glucocorticoid-inducible genes and redundancy ...... 167 6.4 A possible role for transrepression in the dexamethasone-induced repression of inflammatory genes ...... 172 6.5 Implication for glucocorticoid therapy and new drug design ...... 174 6.6 Overall conclusion ...... 179

REFERENCES ...... 180

APPENDIX A: ANTIBODIES, SIRNA, PCR AND CHIP PRIMERS ...... 214

APPENDIX B: COPYRIGHT PERMISSION ...... 221

ix List of Tables

Table 3.1 Effect of MAPK inhibitors on cell viability...... 70 Table 3.2 Densitometry analysis for the effect of LMNA targeting siRNA (LsiRNA) on MAPK phosphorylation...... 76 Table 3.3 Effect of LMNA targeting siRNA (LsiRNA) on inflammatory gene expression...... 78 Table 3.4 Effect of LMNA targeting siRNA (LsiRNA) on IL1B-induced inflammatory protein expression...... 83 Table A1. siRNA sequence for knockdown studies ...... 214 Table A2. Antibodies used for western blot analysis ...... 215 Table A3. Primers used for PCR analysis ...... 216 Table A4. Primers used for ChIP PCR analysis ...... 220

x List of Figures

Figure 1.1 MAPK signalling cascade...... 17

Figure 1.2 Schematic representation of the classical, or canonical, NF-B signalling pathway. . 24

Figure 1.3. Transcriptional control by signalling...... 33

Figure 3.1. Effect of dexamethasone and IL1B on MAPK phosphorylation and DUSP1 expression...... 61

Figure 3.2. Effect of dexamethasone and IL1B on inflammatory gene expression...... 63

Figure 3.3 Effect of JNK inhibitor 8 on IL1B-induced JUN phosphorylation...... 65

Figure 3.4 Effect of MAPK inhibitors on the mRNA expression of inflammatory genes...... 66

Figure 3.5 Effect of MAPK inhibitors on the protein expression of inflammatory genes...... 67

Figure 3.6 Effect of MAPK inhibitors on IL1B-induced inflammatory gene expression...... 69

Figure 3.7 Effect of DUSP1 over-expression on IL1B-induced MAPK phosphorylation and inflammatory gene expression...... 71

Figure 3.8 Effect of LMNA- and DUSP1-targeting siRNAs on IL1B and IL1B plus dexamethasone-induced DUSP1 expression...... 73

Figure 3.9 Effect of LMNA- and DUSP1-targeting siRNA on IL1B-induced MAPK phosphorylation...... 75

Figure 3.10 Effect of DUSP1 targeting siRNAs on IL1B-induced inflammatory gene mRNA expression...... 79

Figure 3.11 Effect of DUSP1 targeting siRNA on IL1B-induced inflammatory gene protein release/expression...... 84

Figure 4.1 Enhanced inflammatory gene expression by IL1B: Feedback control by DUSP1 and feed-forward control by ZFP36...... 93

Figure 4.2 Characterization of TNF expression in the presence of IL1B and dexamethasone. ... 96

Figure 4.3 Analysis of DUSP1 and ZFP36 expression in the presence of IL1B and dexamethasone...... 98

Figure 4.4 Analysis of DUSP1 and ZFP36 expression in the presence of IL1B and MAPK inhibitors...... 100

xi Figure 4.5 Analysis of DUSP1 and ZFP36 expression in primary human bronchial epithelial cells...... 101

Figure 4.6 Analysis of TNF expression in primary human bronchial epithelial cells...... 103

Figure 4.7 Comparison of CT values for TNF expression in A549 and HBE cells generated by RT-PCR...... 104

Figure 4.8 Effect of LMNA siRNA on ZFP36 and TNF expression...... 104

Figure 4.9 Effect of ZFP36 siRNA on IL1B-induced TNF mRNA expression...... 106

Figure 4.10 Effect of DUSP1 over-expression on ZFP36 and TNF mRNA expression...... 108

Figure 4.11 Effect of MAPK inhibitors on TNF expression...... 110

Figure 4.12 Effect of DUSP1 knock-down on ZFP36 and TNF expression...... 112

Figure 4.13 Role of ZFP36 in the enhanced loss of TNF mRNA following DUSP1 knock- down...... 115

Figure 4.14 Effect of MAPK inhibitors and DUSP1 over-expression in the regulation of TNF protein expression...... 117

Figure 4.15 Effect of LMNA siRNA on TNF protein expression...... 117

Figure 4.16 Effect of DUSP1- and ZFP36- targeting siRNAs in the regulation of TNF protein expression...... 118

Figure 4.17 Regulation of TNF gene expression following DUSP1 silencing and glucocorticoid treatment...... 126

Figure 5.1 Effect of DUSP1 over-expression, IL1B and IRF1-targeting siRNA on inflammatory gene expression...... 130

Figure 5.2 Effect of DUSP1 over-expression on IL1B-induced IRF1 expression...... 134

Figure 5.3 Effect of MAPK inhibitors on IRF1 expression...... 136

Figure 5.4 Effect of MAPK inhibitors on IRF1 protein stability...... 139

Figure 5.5 Characterization of IRF1 expression in the presence of IL1B and dexamethasone. . 141

Figure 5.6 Effect of PS1145 on IL1B-induced IRF1 expression...... 143

Figure 5.7 Effect of dexamethasone on IRF1-dependent gene expression...... 145

xii Figure 5.8 Effect of DUSP1-targeting siRNA on IL1B-induced MAPKs and IRF1 expression...... 146

Figure 5.9 Effect of DUSP1-targeting siRNA and MAPK inhibitors on IRF1-dependent gene expression...... 148

Figure 5.10 Effect of IRF1-targeting siRNA on IL1B-induced inflammatory gene expression. 150

Figure 5.11 Characterization of IRF1 occupancy on inflammatory loci in the presence of IL1B and dexamethasone...... 151

Figure 5.12 Effect of PS1145 on IL1B-induced IRF1-dependent inflammatory gene mRNA expression...... 156

Figure 6.1 Regulation of IRF1 and IRF1-dependent late-phase gene expression by IL1B following DUSP1 silencing and glucocorticoid treatment...... 167

xiii

List of Abbreviations

Official Organization (HUGO) committee gene symbols have been used for all genes and gene products.

[Ca2+]i Intracellular Ca2+

ACT D Actinomycin D

ACTH Adrenocorticotropic hormone

Ad Adenovirus

AD5 Adenoviral serotype 5

AF Activation function

AHR Airway hyper-responsiveness

AP-1 Activator protein-1

ARE AU-rich elements

ASK Activation of S-phase kinase

ASM Airway smooth muscle

ATP Adenosine triphosphate

AU Adenine and uridine

AUF1 Adenine-uridine-rich element RNA-binding factor-1

BAL Bronchoalveolar-lavage

BEBM Bronchial epithelial cell basal medium BP Base pairs

BSA Bovine serum albumin

C/EBP CAAT/Enhancer binding protein cAMP Cyclic adenosine monophosphate

CCL (C-C Motif) ligand

xiv CDKN1C/p57KIP2 Cyclin-dependent kinase inhibitor 1C

ChIP Chromatin Immunoprecipitation

CHX Cycloheximide

COX-2 Cyclooxygenase (PTGS2)

CRH Corticotropin releasing hormone

CXCL10 ΙFN-γ-induced chemokine IFN-γ-inducible protein 10

DBD DNA binding domain

Dex Dexamethasone

Dexras Dexamethasone-induced Ras-related protein

DMEM Dulbecco’s modified Eagle’s medium

DoK Downstream of tyrosine kinase ds Double stranded

DSIF DRB (5,6-dichloro-1-β-d-ribofuranosylbenzimidazole) sensitivity-inducing factor

DUSP Dual specificity protein phosphatases

ECL Enhanced chemiluminescence

EDTA Ethylenediaminetetraacetic acid

EGR Early growth response

ERK Extracellular signal regulated kinase

FCS Fetal calf serum

FcR1 High affinity IgE receptor

FEV Forced expiratory volume

FKBP FK506 binding protein

GAPDH Glyceraldehyde-3-phosphate dehydrogenase

xv GFP Green fluorescent protein

GILZ Glucocorticoid-induced (TSC22D3)

GINA Global Initiative for Asthma

GM-CSF Granulocyte-macrophage colony-stimulating factor

GPCR G-protein coupled receptor family

GR Glucocorticoid receptor (NR3C1)

GRE Glucocorticoid response element

GRIP Glucocorticoid receptor interacting protein

H2A Histone 2A

H2B Histone 2B

H3 Histone 3

H4 Histone 4

HAT Histone acetylase

HBE Human bronchial epithelial

HBSS Hank’s balanced salt solution

HDAC Histone deacetylase

HPA Hypothalamic-pituitary-adrenal

HRV Human rhinovirus

HSP Heat shock protein

HuR Human antigen R

ICAM Intracellular adhesion molecule

ICS Inhaled corticosteroids

IFN Interferon

IgE Immunoglobulin E

xvi IKK IBkinase

IL Interleukin iNOS Inducible nitric oxide synthase

IP Immunoprecipitation

IRF Interferon regulatory factor

ISRE Interferon-stimulated response element

IB Inhibitor of B (NFKBIA)

JAK Janus kinase

JNK c-Jun amino terminal kinase

JNK-IN-8 JNK inhibitor 8

KHSRP K-homology domain splicing regulatory protein

LABA Long-acting 2-adrenoceptor agonist

LPS Lipopolysaccharide

MAPK Mitogen activated protein kinase

MHC Major histocompatability complex

MK2 MAPK-activated protein kinases

MKK MAPK kinases

MKKK MAPK kinase kinases

MLCK Myosin light chain kinase

MLK Mixed lineage kinases

MMLV Moloney Murine Leukemia Virus

MOI Multiplicity of infection

MTT 3-(4, 5-dimethylthiazol-2-yl)-2, 5-diphenyltetrazolium bromide

xvii NELF Negative elongation factor

NEMO NF-B essential modulator

NES Nuclear export sequence

NF-AT Nuclear factor of activated T-cells

NFKBIA Inhibitor of B (IB) 

NF-B Nuclear factor--light-chain enhancer of activated B cells nGRE Negative GRE

NLS Nuclear localisation sequence

NP40 Nonidet P-40

NR

NR3C1 Nuclear receptor superfamily 3, group C, member 1

ODU Optical density units

PBMCs Peripheral blood mononuclear cells

PBS Phosphate-buffered saline

PDE Phosphodiesterase

PEF Peak expiratory flow

PHA Phytohemagglutinin

PI3K Phosphoinositide 3-kinase

PIC Protease inhibitor cocktail

PKA Protein kinase A

PMA Phorbol myristate acetate

Pol II RNA polymerase II

P-TEFb Positive transcription elongation factor b

PTPs Protein tyrosine phosphatases

xviii RGS Regulator of G-protein signaling

RIPA Radioimmunoprecipitation assay

RT-PCR Real-time polymerase chain reaction

RV Rhinovirus

S Serine

SABA Short-acting 2-adrenoceptor agonist

SAPK1 Stress-activated protein kinase 1

SEGRA Selective GR/NR3C1 agonist siRNA Short interfering RNA

SLAP Src-like adaptor protein snRNAs Small nuclear RNAs

SRC Steroid receptor coactivator

STAT Signal transducers and activators of transcription

SV Simian virus

T Threonine

TACE TNF-converting enzyme

TAK1 Transforming growth factor-β-activated kinase 1

TBP TATA-binding protein

TH T-helper

TLR Toll-like receptor

TNF Tumor necrosis factor α

Tregs T regulatory cells

TSLP Thymic stromal lymphopoietin

TTP Tristetraprolin

xix UBD Ubiquitin D

UTR Untranslated region V Volts

VCAM Vascular cell adhesion molecule

VEGF Vascular endothelial growth factor

ZFP protein

xx Chapter 1: Introduction

Glucocorticoids regulate a variety of processes, including metabolic homeostasis, cell proliferation, immune and inflammatory responses. Due to their potent anti-inflammatory effects, glucocorticoids are used in the treatment of autoimmune, inflammatory and allergic disorders, including asthma, rheumatoid arthritis, ulcerative colitis, inflammatory bowel disease and allergic rhinitis (1). This thesis investigates the molecular mechanisms of inflammatory gene repression by glucocorticoids in the context of the cells relevant to airways inflammation and asthma. In addition, poor clinical responses to glucocorticoids by asthmatics who smoke, or have viral infections, illustrates the need to understand the detailed molecular mechanism(s) of action in order to promote the development of novel glucocorticoid ligands or add-on therapies with an enhanced anti-inflammatory profile.

1.1 Asthma

According to the Global Initiative for Asthma (GINA) 2009, “asthma is a chronic inflammatory disorder of the airways in which many inflammatory cells and cellular events play a role”.

Clinically, the chronic inflammation in asthma is typically associated with symptoms such as wheezing, cough, breathlessness and a variable airflow obstruction that are attributed to non- specific airway hyper-responsiveness (AHR), reversible airway bronchoconstriction and structural changes to the airways (airway remodelling) (GINA, 2009). As per disease severity, asthma can be classified into four categories: intermittent, mild persistent, moderate persistent and severe persistent (2). A number of factors, including viral respiratory infections, host responses, environmental pollutants and tobacco smoke, can not only trigger an asthma attack, but may also influence the risk of initially developing asthma (GINA, 2009). Asthma is one of the most common chronic disease affecting children and young adults (3, 4). However, the prevalence of asthma is

1 more common in children than adults, and susceptibility to developing asthma in childhood is believed be determined by the events that occur in the first three to five years of life (Public Health

Agency of Canada). In addition, exposure to airborne allergens, environmental toxins and tobacco smoke along with frequent respiratory infections in early life are the strongest risk factors associated with the development of childhood asthma (Public Health Agency of Canada).

Although asthma has both genetic and environmental components, a family history of allergic and atopic diseases, such as hay fever and eczema can also influence the risk of developing asthma in children (5).

1.1.1 Burden of asthma

The prevalence, morbidity, mortality and economic burden associated with asthma is increasing in Canada and the rest of the world (6, 7). It is estimated that around 235 million people are affected by asthma worldwide (WHO). Although asthma is more common in high-income countries, some low- and middle-income countries also have high levels of asthma prevalence (8). If this current trend continues, it is projected that by 2025 there will be a 100 million additional asthmatics globally (2). In Canada alone, over 3.2 million of the population have been diagnosed with asthma, and is expected to rise to over 3.9 million by 2030 (statistics Canada, Public Health Agency of

Canada). Even though the number of hospitalizations have been declining over the past 20 years, in 2004, a staggering 10% of the hospitalizations among children was due to asthma (GINA, 2004).

This increasing number of hospitalizations reflects an increase in severe asthma and poor disease management (8). Although the mortality rates have declined significantly in the world, around

180,000 deaths are attributed to asthma every year (WHO). The majority of asthma-related deaths occur in those aged ≥45 years (8). Despite the availability of more efficacious medications, asthma

2 still affects up to 20% of children and causes ~500 deaths per year in Canada (Statistics Canada,

Public Health Agency of Canada).

The cost of asthma creates a considerable economic burden and the patients with severe asthma are responsible for approximately 50% of all direct and indirect costs (GINA, 2004). Direct costs of asthma treatment include hospitalization and medication, and were estimated to exceed $700 million in Canada in 2000 (Public Health Agency of Canada). In view of this, it has been suggested that because unplanned or emergency hospitalization is more expensive than planned treatments, adequate asthma prevention could in fact decrease the overall financial burden of the disease

(GINA, 2009). There are also significant indirect, socio-economic costs associated with asthma, including premature deaths and long-term disability (GINA, 2009). In Canada, these indirect costs were estimated to total $840 million in 2000 (Public Health Agency of Canada) and includes the financial burden related to school absenteeism and days lost from work (GINA, 2009). Asthma accounts for approximately 1% of all disability-adjusted life years, which reflects the high prevalence and severity of the disease (GINA, 2009).

1.1.2 Asthma pathogenesis

The pathophysiology of asthma is complex, but typically involves allergic inflammation of the airways (9). Allergic inflammation in asthma involves increased infiltration of mast cells, macrophages and in particular eosinophils into the airways (10). However, in severe asthma neutrophils are also recruited into the airways (11). Recruitment of inflammatory cells brings about pathological changes through the release of inflammatory mediators, which in turn activate signalling pathways to enhance the expression of multiple inflammatory cytokines, chemokines, adhesion molecules, growth factors, inflammatory enzymes and receptors (10). In particular, the

3 inflammatory response in allergic asthma involves increased synthesis and release of cytokines, such as interleukin (IL) 4, IL5, IL9 and IL13, which, together, are responsible for IgE production from B-cells and eosinophilic inflammation (9, 12).

The allergic response in asthma involves the activation of resident mast cells and dendritic cells by allergen, consequently resulting in the activation of IgE mediated responses (13). These dendritic cells are generally separated from the airway lumen by a layer of epithelium. In order to access the dendritic cells, inhaled allergen either disrupts epithelial tight junctions or increases the permeability of the epithelium (14). For example, the house dust mite allergen, Der p1, possesses intrinsic protease activity and, by cleaving tight junction proteins directly, can access the underlying dendritic cells (15). Allergen-activated dendritic cells migrate to the draining lymph

+ + nodes and present the processed antigen to naïve CD4 TH cells (10). These CD4 TH cells are subsequently activated by binding of the T cell receptor and CD28 on the T cell surface to the antigen-major histocompatability (MHC) complex and CD80/86 on the dendritic cell (16). This co-stimulation serves as a verification signal for T cells, as without it, T cell become anergic, and irresponsive to antigen (16). This mechanism is essential to prevent self-antigen-mediated immune responses (16).

Upon co-stimulation, T cells regulate the production of IL2 and IL4, which induces TH cell proliferation and initiates the polarisation of T cells to a TH2 phenotype respectively (17). TH2 cells predominate in the asthmatic airways and by releasing a number of cytokines, such as IL4,

IL5, IL9 and IL13, TH2 cells play an important role in allergic inflammation associated with asthma (9). The TH2 cells-derived IL5 and IL9 are further involved in the differentiation of eosinophils and mast cells respectively, whereas IL4 and IL13, by binding to their receptors on B

4 cells, initiate the synthesis and release of IgE antibodies (9). These IgE antibodies circulate briefly in the blood before binding to high-affinity IgE receptors, FcRI, on the surface of mast cells (9).

Upon re-exposure, allergen binds to IgE-FcRI and activates the mast cells via the cross-linking of IgE-FcRI (18). Activated mast cells, by releasing multiple inflammatory mediators, contribute to both acute and chronic allergic inflammation (10). The acute phase of allergic inflammation, commonly referred to as the ‘early phase’ of allergic reaction, typically occurs within minutes, or even seconds, following allergen exposure and usually resolves within an hour (10). Mast cell- derived bronchoconstrictors, such as histamine, cysteinyl leukotrienes and prostaglandin D2, initiate the ‘acute/early phase’ of the allergic response (9, 10), and is often associated with mucosal oedema, mucus production and contraction of the airway smooth muscle (10).

The ‘early phase’ of allergic responses are followed by a more sustained chronic inflammation, known as the ‘late phase’ of the allergic response (18). This late allergic response involves increased infiltration of TH2 cells, eosinophils and basophils into the airways via mast cell-derived mediators, including leukotrienes, cytokines and chemokines (9, 18). The ‘late phase’ of the allergic response usually peaks around 8-12 hours after allergen exposure and typically involves increased bronchoconstriction, sustained edema, mucus overproduction and AHR to bronchoconstrictors, such as histamine and methacholine (18). Chemokines, such as chemokine

(C-C Motif) ligand (CCL) 20 and CCL17, recruit TH2 cells into the airways (19). These TH2 cells further enhance the synthesis and release of TH2 cytokines, such as IL4, IL9, IL13, IL3, IL5 and granulocyte-macrophage colony-stimulating factor (GM-CSF; CSF2) (10). Enhanced IL5 stimulates the release of eosinophils and prolongs their survival (9). To participate in the allergic inflammatory response, eosinophils migrate from the blood circulation into the airways up

5 concentration gradients of chemokines such as CCL2, CCL3, CCL5 and CCL11 (20). Upon maturation, eosinophils degranulate and release inflammatory proteins, such as major basic protein, eosinophil neurotoxin, eosinophil cationic protein and eosinophil peroxidase (21).

Eosinophil-derived major basic protein directly damages the airway epithelium, enhances bronchial responsiveness, and initiates degranulation of basophils and mast cells (21). In contrast, cysteinyl leukotriene C4 released by eosinophils contracts airway smooth muscle and increases vascular permeability (22). Even though phenotypically asthmatic inflammation is eosinophilic, a number of studies indicate that virus-induced asthma exacerbations are associated with increased production of the neutrophilic chemokine, IL8, and involves neutrophilic inflammation (23, 24).

The expression of the neutrophil-derived protease, neutrophil elastase, which stimulates mucus production, is also enhanced in neutrophilic asthma and is therefore likely to play a key role in asthma pathogenesis (24).

The allergic inflammation in asthma arises from a relative imbalance between TH1 and TH2 cells

(9). The TH1 cell enhances the production of TH1 cytokine, interferon-γ (ΙFN-γ), which inhibits the synthesis of IgE and the differentiation of precursor cells to TH2 cells (10). Thus, in the absence of ΙFN-γ there may be an enhanced release of TH2 cytokines, which may further induce allergic inflammation (10). In this regard, studies also indicate that TH1 cytokines, ΙFN-γ and IL12, by antagonizing the TH2 responses, inhibit allergic inflammation associated with asthma (25, 26).

Therefore, the TH1 response may be considered protective for asthma, whereas the TH2 response is often linked with severe allergic inflammation (27). However, in vivo studies from asthma show conflicting results and do not fully support this hypothesis (10). For example, the serum levels of

ΙFN-γ were elevated in severe asthmatic patients (28). Similarly, ΙFN-γ levels were also increased

6 in supernatants from phytohemagglutinin (PHA)/phorbol myristate acetate (PMA)-stimulated bronchoalveolar-lavage (BAL)-fluid cell cultures derived from asthma patients (29). Enhanced levels of IFN-γ, by increasing the expression of CD69, HLA-DR and intercellular adhesion molecule (ICAM) 1, contributes to the activation of eosinophils, and thus, is likely to enhance eosinophilic inflammation (10, 30). Therefore, these data point to the possibility that both TH1 and

TH2 cells might contribute to airway inflammation in asthma. Besides TH1-TH2 imbalance, other

TH cells such as TH17, TH22, TH9 and T regulatory cells (Tregs) also play a key role in asthma pathogenesis (31).

If the inflammation in asthma is uncontrolled, it may further lead to structural changes of the airways known as airway remodelling (11). Characteristic changes include airway smooth muscle hypertrophy, basement membrane thickening and angiogenesis (32). In addition, goblet cell metaplasia along with hyperplasia and enlargement of submucosal glands stimulate the secretion of mucus into the airway lumen (10). Airway remodelling may contribute to airflow obstruction and AHR (11). Until recently, it was proposed that airway remodeling may develop late in the disease process as a consequence of uncontrolled or persistent inflammation (33). However, recent findings show that the symptoms associated with airway remodelling, such as basement membrane thickening and epithelial cell disruptions, are detectable very early in the disease pathogenesis especially in children with or without asthma diagnosis (33, 34). Airway remodelling is thereby considered a crucial component of childhood asthma (34). To this end, understanding the underlying mechanisms responsible for structural changes may be helpful in the clinical management of childhood asthma (33).

7 1.1.3 Exacerbation of asthma

Asthma exacerbations can be defined as “….episodes characterized by progressive increase in shortness of breath, cough, wheezing and chest tightness, or some combination of these, and increased airflow obstruction that is manifested by reduction in measurements of lung function such as peak expiratory flow (PEF)” (35). Exacerbations of asthma accelerate disease progression, lower quality of life of asthmatics, increase morbidity, mortality and economic burden of the disease (36, 37). A number of risk factors, including allergen, pollutant and bacterial infections can trigger asthma exacerbation (38). However, infection with respiratory viruses, including human rhinovirus (HRV), influenza virus and respiratory syncytial virus, are major risk factors associated with exacerbation of asthma in both adults and children (39, 40). Studies have indicated that infection with HRV alone is responsible for 60% of asthma exacerbations in children and about 45% of asthma exacerbations in adults (41, 42). Viral-induced asthma exacerbations are mainly characterised by neutrophilic airway inflammation, mucus hypersecretion and bronchial hyper-responsiveness (43). In addition, there is also an increased secretion of a number of cytokines and chemokines, including IL8, IL6, IL1, CCL5 and CXCL10 (44-47), and is mainly attributed to the increased activation of the pro-inflammatory transcription factor, nuclear factor-

-light-chain enhancer of activated B cells (NF-B) (48). Enhanced release of the neutrophilic chemokine, IL8, contribute to neutrophilic inflammation, whereas CXCL10 and CCL5 are responsible for lymphocytic and eosinophilic inflammation respectively (48). Viral infections are also coupled with impaired innate and adaptive immune responses of asthmatics (49, 50). For example, bronchial epithelial cells obtained from asthmatics exhibit enhanced viral replication leading to increased epithelial cell damage and decreased innate immune responses (50). It has also been proposed that impaired production of IFNs during viral infections not only decrease

8 macrophage responses to virus infection, but may also reduce anti-viral immune responses (35,

51). In addition, the synthesis of both type I IFN- and type III IFN- is also deficient in asthmatics compared to healthy individuals following HRV infection and may further correlate with the severity of HRV-induced asthma exacerbations (49, 52, 53). Furthermore, asthmatics who smoke also have a higher frequency of exacerbation with an increased risk of hospitalization during exacerbations (54, 55).

1.1.4 Pulmonary epithelium and its role in asthma

A continuous layer of epithelial cells lines the airways from the trachea to the alveoli, and is the interface between the host and the environment (56). Morphologically, while the airway epithelium in the large airways is pseudostratified, the epithelium in the small airways is columnar and cuboidal (57, 58). The epithelium in the airways is comprised of a variety of cell types, including ciliated, columnar, undifferentiated, secretory, and basal cells. In contrast, the epithelium in the alveoli is composed of type I and type II alveolar cells (57, 58). The ciliated epithelial cells together with basal cells forms a pseudostratified epithelium and play a major role in the removal of foreign particles trapped in the mucus lining the upper airways (mucociliary clearance) (58).

The type I alveolar epithelial cells are thin, flat, squamous cells, covering 95% of the alveolar surface and play a vital role in the gas exchange (59). On the other hand, type II alveolar epithelial cells are cuboidal in appearance and cover 5-10% of the alveolar epithelium. The type II cells secrete surfactant, which reduces alveolar surface tension, and by acting as progenitor cells, the type II cells can repopulate type I cells in the alveolar epithelium (60).

Airway epithelial cells not only act as a first line of defense against inhaled pathogens, allergens and other foreign particles but also possess a wide variety of functions, including repair,

9 regeneration and the regulation of structural and immune cells (57, 58). Airway epithelial cells, by releasing pro-inflammatory cytokines, chemokines, and inflammatory peptides, including IL8,

CXCL5, CXCL10, CCL5, IL1, IL6 and others, contribute to the inflammatory responses in asthma

(61, 62). Epithelial cell-mediated secretion of IL8 and CCL5 increases the recruitment of neutrophils and eosinophils respectively into the airways (9), which in turn, by releasing proteases and cytotoxic mediators in the airways, damages the epithelial cell layer (21). Disruption of airway epithelial integrity is often regarded as an important mechanism in the development of AHR (61).

Enhanced AHR could be attributed to the possibility that the loss of epithelial barrier may allow inhaled allergen to access the mucosa and underlying immune cells. In addition, epithelial cells release large amounts of the cytokine, thymic stromal lymphopoietin (TSLP) that orients dendritic cells towards the TH2 response and triggers allergic inflammation (63, 64). Furthermore, the stem cell factor, produced by airway epithelial cells is involved in the recruitment of mucosal mast cells to the airway surface. These mast cells release bronchoconstrictors, such as histamine and leukotrienes, and further produce bronchoconstriction (9, 65). However, contrary to their role in airway inflammation, epithelial cells, by mediating the release of antimicrobial, complement and other factors, may also contribute to innate immune and host defense responses (66, 67). Equally, airway epithelial cells also produce enzymes, permeabilizing peptides, collectins and protease inhibitors that are involved in the killing of microorganisms, and thereby airway epithelium provides a first line of defense against pathogenic insults (61).

As mentioned above, airway epithelial cells release a number of inflammatory mediators that are involved in the pathogenesis of asthma. In addition, airway epithelium is also the first point of contact for asthma therapy, including inhaled glucocorticoids. Thus, examining the repressive

10 effects of glucocorticoids on airway epithelial cell-derived inflammatory mediators may assist into understanding the molecular mechanism of repression of inflammatory genes by glucocorticoids

(61, 62).

1.1.5 Current asthma therapy

The aim of asthma treatment is not only to reduce or combat the underlying inflammatory process in order to reduce disease progression, improve lung function, reduce symptoms, and exacerbations but also to minimize the side effects associated with asthma therapy (GINA, 2009).

Treatment of asthma consists of a step-wise approach with administration of a reliever treatment, such as short-acting 2-adrenoceptor agonist (SABA), taken “as-needed” for mild asthmatics

(GINA, 2009). SABAs are used as a bronchodilators to relieve symptoms, such as wheezing, chest tightness and cough (GINA, 2009) (68). If the symptoms are not controlled and worsen or become more frequent with reliever medication, addition of one or more preventer medications, including inhaled corticosteroids (ICS), is recommended (GINA, 2009).

ICS is currently the most effective anti-inflammatory therapy available for the treatment of asthma

(68). Despite the widespread use of ICS, many patients with severe asthma and asthmatics who smoke show poor responsiveness to ICS due to severe inflammation. These patients require high, often oral, doses of glucocorticoid over prolonged periods to combat inflammation (69, 70). In such cases, therapeutic usefulness is compromised by systemic contraindications that are, in part, attributed to the metabolic effects of glucocorticoids and to the effects on the hypothalamic- pituitary-adrenal (HPA) axis (71, 72). These side-effects include osteoporosis, muscle wasting, stimulation of gluconeogenesis, increased blood glucose levels, diabetes, cataracts, weight gain, and impaired skeletal muscle growth in children (72, 73). Importantly, these contraindications

11 reduce the usefulness of ICS for the treatment of severe asthma. Moreover, ICSs are also less effective or have no effect on the decline in lung function and they often fail to prevent the progression of asthma in young children (74). Due to this, the dose of ICS may be increased or long-acting 2-adrenoceptor agonist (LABA) may be added to therapy (GINA, 2009). Because long term use of high dose ICS is coupled with increased risk of side-effects (see above), the addition of LABA is preferred over increasing the dose of ICS (GINA, 2009).

LABAs, such as salmeterol and formoterol, act by binding to cell-surface 2-adrenoceptors, which are members of the 7-transmembrane, G-protein coupled receptor (GPCR) family. Upon binding to the 2-receptor, LABA triggers dissociation of Gssubunit of GPCR from the / dimer and further activates adenylate cyclase, which catalyses the conversion of adenosine triphosphate

(ATP) to 3', 5'-cyclic adenosine monophosphate (cAMP), and thereby enhances the intracellular concentration of this second messenger. Consequently, this leads to the activation of the cAMP- dependent pathway where cAMP activates an enzyme protein kinase A (PKA). Activated PKA further phosphorylates a number of downstream targets and produces bronchodilation (75-78). For example, PKA-mediated phosphorylation of myosin light chain kinase (MLCK) reduces the ability of MLCK to phosphorylate myosin and consequently inhibits smooth muscle contraction (76). In addition, cAMP by inhibiting the release of free calcium (Ca2+) from intracellular stores reduces the level of intracellular Ca2+ ([Ca2+]i). Since, increased [Ca2+]i is responsible for the contraction of airway smooth muscle, cAMP-mediated attenuation of [Ca2+]i results in bronchodilation (76).

Together, these mechanisms attenuate smooth muscle contraction to provide relief from bronchoconstriction.

12 Current asthma treatment includes a combination therapy incorporating ICS and LABA to reduce both airway inflammation and bronchoconstriction. Inhalers containing a combination of ICS plus

LABA (GSK: Seretide/Advair (salmeterol/fluticasone propionate) and AstraZeneca: Symbicort

(budesonide/formoterol)) are increasingly prescribed due to their higher clinical efficacy (79-81).

In 1994, Greening et al. first observed the effectiveness of LABA/ICS combination in improving the symptoms and lung functions following the addition of salmeterol to ICS (82). Since then, many clinical studies have demonstrated superior control of asthma with LABA/ICS combination therapy, showing synergistic effect on bronchodilation as well as reduction in the symptoms, improvement in the quality of life and lower frequency of exacerbations (81). In addition, work from Woolcock et al., Pauwels et al. and Wilding et al. show that the addition of LABA to a standard dose of ICS is more effective clinically than increasing the dose of ICS (75). A number of other in vivo and in vitro data are also available explaining the possible mechanisms mediating the increased effectiveness and clinical efficacy of combination therapy (75). In this context, glucocorticoids augment the density of 2-receptor (83), attenuate functional desensitisation of the receptor (84) and increase both Gs expression and coupling to adenylyl cyclase (85).

Glucocorticoids also prevent cytokine (for example IL1B)-mediated hypo-responsiveness of human airway smooth muscle (ASM) cells to 2-receptors (86, 87). In contrast, the mechanism by which LABA increases the activity of glucocorticoids is largely unknown. However, LABA increases the clinical efficacy of ICS to a level that cannot be achieved by ICS alone (75, 88).

Glucocorticoid-mediated histone acetylation (89) and anti-inflammatory gene expression is also enhanced by LABA (75). The translocation of glucocorticoid receptor (GR; NR3C1) into the nucleus upon activation by glucocorticoids is also potentiated by LABA (90).

13 Other preventer medications, such as leukotriene modifiers and/or theophylline, are also recommended as a maintenance therapy for symptomatic control of asthma (GINA, 2009).

Leukotriene modifiers, such as 5’-lipoxygenase inhibitors and leukotriene receptor antagonists, act by blocking the production and receptor binding respectively, of leukotrienes (91). Leukotriene receptor antagonists are particularly used to reverse the bronchoconstriction induced by exercise or smoking (54, 91). In contrast, theophylline is a non-selective inhibitor of phosphodiesterase

(PDE), that increases the intra-cellular concentration of cAMP and produces bronchodilation (71).

Theophylline is generally used to reverse bronchoconstriction in asthmatics who are smokers (54).

However, the utility of theophylline is limited by its very low therapeutic index (68).

1.2 Control of inflammatory gene expression

Inflammatory gene expression can be controlled transcriptionally or post-transcriptionally by processes, such as mRNA stability, translation and protein stability, via the activation of one or more inflammatory transcription factors and/or cellular signalling mechanisms.

1.2.1 Mitogen activated protein kinase (MAPK) pathways

The MAPK pathways are evolutionarily conserved protein kinase cascades, activated by extracellular stimuli such as UV radiation, pro-inflammatory cytokines, including TNF and IL1B, oxidative stress, T- and B-cell receptor activation, and other mitogens (92, 93). Extracellular signal regulated kinase (ERK) 1/2, also known as p44/p42, was the first mammalian MAPK to be characterised (94, 95). Later on, other MAPKs including the c-Jun amino terminal kinase (JNK), also known as stress-activated protein kinase 1 (SAPK1), and p38 MAPK, also known as SAPK2-

4, were characterised (96). The signalling cascade begins with the activation of MAPK kinase kinase (MKKK) at the cell surface receptor by small GTP binding proteins, GTPases (e.g. Ras,

14 cdc42), or other protein kinases (97). These MKKKs phosphorylate and activate a number of

MAPK kinases (MKKs or MEKs), which in turn dual phosphorylate the serine/threonine and tyrosine residues of the MAPKs, ERK, p38 and JNK at a T-x-Y motif, leading to MAPK activation

(Fig. 1.1) (92, 98). Thus, ERK is activated by MKK1/MKK2, p38 by MKK3/MKK6, and JNK by

MKK4/MKK7 (96). In addition, other MKKKs, including MKKK1-4, c-raf, mixed lineage kinases, activation of S-phase kinase (ASK) 1 and 2, transforming growth factor (TGF)-β-activated kinase 1 (TAK1) and tumour progression locus 2 can also activate MAPKs (99). Activated MAPKs by phosphorylating various substrates, including mRNA destabilising proteins, cytoskeletal proteins and transcription factors, regulate the expression of specific genes within the target cells

(94, 100). In this regard, MAPKs exert a number of cellular responses, including proliferation, differentiation, survival, apoptosis, embryogenesis, and gene expression, and some of these responses are mediated via the phosphorylation of transcriptional factors such as c-Jun, ATF-2, , Elk-1, NFAT, NF-B and/or AP-1 (94, 95, 97, 101, 102). In addition, p38 MAPK is involved in the phosphorylation of multiple kinases, including MAPK-activated protein kinases (MK2 &

MK3), MSK1, MSK2, MNK1 and MNK2 (102). Conversely, JNK-mediated phosphorylation of the E3 ubiquitin ligase, Itch, leads to the functional activation of Itch and consequently enhanced degradation of c-Jun and JunB (103).

Since MAPKs are activated by phosphorylation of both the threonine and tyrosine residues, dephosphorylation of either of residue is sufficient for complete inactivation (104). This can be achieved by tyrosine-specific protein tyrosine phosphatases (PTPs), serine/threonine-specific phosphatases, or by dual specificity (threonine/tyrosine) protein phosphatases (DUSPs) (104).

Serine/threonine phosphatases, such as PP2A, by dephosphorylating the threonine residue,

15 inactivates both MAPK and MKK (104, 105). In this context, PP2A by dephosphorylating ERK and c-Src, an upstream kinase of ERK signalling, inhibits ERK MAPK pathway (106). In contrast, the serine/threonine phosphatase, PP2C inactivates p38, JNK and MKKs (105). Similarly,

PTPs, such as PTP-SL, haematopoietic PTP and leukocyte PTP, by dephosphorylating the tyrosine residue, targets ERK1/2 and p38 for inactivation (105, 107).

Finally, DUSPs can dephosphorylate both the threonine and tyrosine residues on the same substrate (104, 105). To date, 9 family members of DUSP have been identified. DUSP1, also commonly referred to as MAPK phosphatase 1, was the first DUSP to be discovered in 1991 (108).

All the DUSPs share a common structure and have a different specificity for one or more MAPKs

(104, 105, 109). Thus, DUSP1 has a specificity for ERK, p38 and JNK, whereas DUSP5 has a specificity for ERK and can only dephosphorylate ERK (109). A number of pro-inflammatory stimuli, including TNF and IL1B, that activate MAPK pathways can also induce the expression of

DUSPs (100). The activity of DUSP proteins can be modulated by various post-translational modifications, including acetylation, phosphorylation, ubiquitination and degradation (109). In this context, phosphorylation increases the stability of DUSPs by attenuating the ubiquitination and degradation of DUSPs (109).

16 Stimulus e.g. TNF or IL1B

GTP-binding protein Rac, Cdc42 Ras

MKKK/MEKK TAK1/ASK1 MKKK1/2 Raf

MKK/MEK MKK3/6 MKK4/7 MKK1/2

MAPK p38///d JNK1/2/3 ERK1/2

Targets MAPKAPK c-Jun c-Fos,Elk-1 NF-B ATF-2

Figure 1.1 MAPK signalling cascade. Schematic diagram of the MAPK signalling pathway for the three major MAPKs p38, ERK and JNK (reviewed from (98, 100, 110)).

In addition, the expression of DUSP proteins is often dependent on the activation of MAPKs and as such, DUSPs, by a negative feedback regulatory loop, control the duration and amplitude of

MAPK signalling (111). Thus, p38 MAPK-mediated expression of DUSP1 is involved in a feedback inhibition of p38 MAPK (112). For further details about the role of DUSP1 in a negative regulation of MAPK signalling and inflammation refer to section 1.3.4.1.

1.2.1.1 Role of MAPK pathways in inflammation and asthma

Since activation of MAPKs is associated with immune and inflammatory responses, respiratory burst activity, chemotaxis, granular exocytosis, adherence, apoptosis, or T-cell mediated differentiation, it is believed that the MAPKs may play an important role in asthmatic

17 inflammation (113). In this context, p38 MAPK, by increasing the expression of the neutrophilic chemokine, IL8, enhances the migration and recruitment of neutrophils into the asthmatic airways, and may thereby contribute to neutrophilic inflammation (114). Furthermore, p38 MAPK by affecting the activity of inflammatory transcription factor, NF-B, regulates the expression of a number of pro-inflammatory cytokines (115). In addition, p38 MAPK also modulates the stability and translation of pro-inflammatory cytokines through the regulation of MNK1 (116) and MK2/3

(117). Moreover, deletion of p38 MAPK in epithelial cells was associated with reduced pro- inflammatory gene expression (118). These observations clearly suggest that the activation of p38

MAPK not only increases asthmatic inflammation, but may also worsen the existing inflammation by augmenting the synthesis and release of pro-inflammatory mediators. Conversely, ERK has been shown to promote the differentiation of naive T cells to a TH2 phenotype, and may thereby contribute to allergic inflammation (119). Since both ERK and p38 MAPKs are involved in the migration of mast cells and eosinophils, their role in the development of asthma is implicated (120,

121). In addition, ERK and JNK, by increasing the expression and phosphorylation of c-Fos and c-Jun, control the activity of inflammatory transcription factor AP-1 (98, 122, 123). Similar to NF-

B, AP-1 regulates the expression of a number of inflammatory and immune mediators, including, but not limited to, IL8, CSF2, IL2, TNF and CCL2, that are involved in the pathogenesis of asthma

(124). Equally, JNK has been shown to enhance the recruitment of eosinophils into the asthmatic airways through enhanced expression of IL1B-induced eosinophilic chemokines, CCL11 and

CCL2, and thus, may participate in eosinophilic inflammation (125). These observations clearly support the role of MAPKs in the inflammatory response associated with asthma. However, contrary to this, inhibition of MAPKs can also contribute to asthmatic inflammation. For example, p38 MAPK inhibitors, or genetic disruption of p38 MAPK pathway, potentiate TH1 responses in

18 macrophages and dendritic cells via the increased production of IL12 (126-128). Since the TH1 response contributes to asthma pathogenesis (see section 1.1.2), inhibition of p38 MAPK could also enhance inflammation in asthma. Similarly, inhibition of MEK1, in human airway epithelial cells, enhances the expression of HRV-induced CXCL10 (129). Because CXCL10 is associated with increased airway inflammation and AHR (130), inhibition of ERK MAPK may also, in part, contribute to asthmatic inflammation. In addition, MKK1-/- T cells, upon T-cell receptor engagement, exhibit attenuated JNK activity and elicit enhanced release of TH2 cytokines, including IL4, IL5 and IL13 (103). Thus, loss of JNK activity could contribute to allergic inflammation through augmented TH2 responses.

1.2.2 Transcription

Transcription plays an important role in the regulation of gene expression and occurs in the nucleus, where DNA is tightly packaged into highly organised nucleoprotein structures known as chromatin (131). The structural unit of chromatin is the nucleosome, which consists of 146 bp of

DNA wrapped around a protein core containing two copies of each of the four histone proteins

H2A, H2B, H3 and H4 (132). Packaging of DNA by histone proteins prevents the access of transcription machinery to DNA and thus, unwinding of DNA from histone is essential for transcription to occur (131, 133). In addition, unwinding of DNA also involves the recruitment of histone modifying enzymes and ATP-dependent chromatin remodelling complexes to DNA which catalyze acetylation, phosphorylation and methylation of histones at N-terminal tails (133).

Acetylation and phosphorylation of histone proteins removes the positive charge on histones, thereby decreasing the interaction of histones with negatively charged DNA (134). As a consequence, DNA becomes unwound from histones, and chromatin is transformed into a more

19 relaxed state that is accessible by the transcriptional apparatus (135). This process is collectively referred to as ‘chromatin remodeling’ (135-137).

For transcriptional activation to occur, chromatin remodelling is followed by a subsequent binding of TATA-binding protein (TBP), RNA polymerase II (Pol II) and basal transcription complexes to a gene promoter to initiate transcription (137). This occurs at the TATA box, identified by a consensus 5’-TATAAA-3’ sequence that is generally found 25-30 bp upstream of the transcription start site (138, 139). Subsequently, through a conformational change, 11-15 bp DNA surrounding the transcription start site is unwound and the active site of Pol II accesses the template strand of the promoter to initiate transcription (140). During the initiation of transcription, a large number of short (3 - 10 nucleotides) transcripts are synthesized and released (140). The next step of transcription involves elongation of the transcript and is associated with hyper-phosphorylation of the C-terminal domain of Pol II (137). Following phosphorylation of C-terminal domain, Pol II dissociates from DRB (5,6-dichloro-1-β-d-ribofuranosylbenzimidazole) sensitivity-inducing factor (DSIF) and negative elongation factor (NELF), thereby gets released from the promoter

(promoter escape) (141). This results in the initiation of the elongation phase of transcription.

Elongation proceeds further with the help of positive transcription elongation factor b (P-TEFb), which facilitates efficient movement of Pol II along the DNA by preventing Pol II pausing (142).

In addition, other factors, such as histone disassembly/reassembly factors, transcription and elongation factor SPT6, enhance transcriptional efficiency by facilitating the movement of Pol II through the chromatin (143).

In order to generate a mature mRNA transcript, pre-mRNA undergoes further processing, including 5’ capping, splicing and generation of the 3’ poly(A) tail (144). Modification of the 5’

20 end of pre-mRNA by addition of 7-methylguanosine is called capping (144), and usually occurs very early during the transcription process, even before the mRNA molecules are finished being transcribed by Pol II (131, 144). The 5’ capping of the mRNA is important to protect the mRNA transcript from degradation by exonucleases (144). The next step of the mRNA processing is the removal of introns by splicing, which is catalysed by a complex of proteins and small nuclear

RNAs (snRNAs), called the spliceosome (131, 144). The spliceosome is recruited to the mRNA via an interaction with the Pol II C-terminal domain (131, 144). Finally, upon reaching the end of a gene, Pol II terminates transcription, the newly synthesized immature transcript gets cleaved, and several hundred adenosine residues constituting the poly(A)tail are added at the 3’ end of the transcript by poly(A) polymerase (144). Subsequently, Pol II dissociates from the DNA and gets recycled to initiate further rounds of transcription (144).

1.2.3 Inflammatory transcription factors

Increased expression of inflammatory mediators is a hallmark of inflammation in asthma. The

DNA-binding proteins, commonly referred to as transcription factors, control the process of transcription via the recruitment of Pol II to a gene promoter and further regulate the expression of inflammatory genes (145). Several inflammatory transcription factors such as NF-B, AP-1,

CAAT/enhancer binding protein (C/EBP), signal transducers and activators of transcription

(STAT) and nuclear factor of activated T-cells (NF-AT) have been implicated in mediating inflammation associated with asthma (145). However, in the context of current project only the roles of NF-B and interferon regulatory factor (IRF) 1 will be discussed.

21 1.2.3.1 NF-B

NF-B is a ubiquitous inflammatory transcription factor, discovered in 1986 via its interaction with an 11- sequence in the immunoglobulin light chain enhancer element (146). NF-B belongs to the Rel family of transcription factors and may be composed of hetero or homodimeric combinations of five different subunits; NF-B1 (p50 and its precursor p105), NF-B2 (p52 and its precursor p100), p65 (RelA), RelB and c-Rel (147). Structurally, NF-B consists of a highly conserved DNA-binding/dimerization domain known as the that regulates the nuclear localisation and DNA binding (148). Furthermore, p105 (NF-B1) and p100 (NF-

B2), contain multiple ankyrin repeats in their C-terminal domain, and are activated via proteolysis, to form the small DNA binding proteins, p50 and p52, respectively (149). In contrast, p65 (RelA), RelB and c-Rel that contain a C-terminal transactivation domain and are therefore capable of activating transcription (149). Conversely, p50 and p52 lack a transactivation domain, and thus, by binding to cRel, RelB, and p65, facilitate the activation of classical or canonical pathway of NF-B activation (150).

A number of stimuli, including pro-inflammatory cytokines, such as TNF and IL1B, mitogens, stress, viruses, double stranded (ds) RNA, phorbol ester, bacteria and bacterial lipopolysaccharide

(LPS) can activate NF-B (151, 152). The activation of NF-B is strongly associated with increased cell survival and inflammatory responses as observed during cancer, chronic inflammatory disorders, including asthma, arthritis and inflammatory bowel disease, and several other neurodegenerative or inflammatory disorders (153). Under resting condition, NF-B is maintained in an inactive form in the cytoplasm through its interaction with an inhibitory protein, inhibitor of B (IB) (NFKBIA) (153). Upon stimulation with an appropriate stimulus, NF-B

22 is activated through a classical (canonical) pathway (Fig. 1.2) (154). This involves the activation of IBkinase (IKK) complex, containing the two catalytic subunits IKK1 (IKK) and IKK2

(IKK), along with a structural subunit referred to as NF-B essential modulator (NEMO; IKK

(155). Among these, IKK2 is essential for the phosphorylation of NFKBIAat serine (S) 32 and

36 residues (151). The phosphorylated NFKBIA is further recognised by ubiquitin ligase machinery, resulting in the poly-ubiquitination and subsequent degradation of NFKBIAby the 26S proteasome. Consequently, p50 and p65 dissociates from NFKBIA and translocates into the nucleus (Fig. 1.2) (151, 155, 156). Once in the nucleus, NF-B binds to specific B sequences, 5’-

GGGRNWYYCC-3’ (R, A/G; N, any nucleotide; W, A/T; Y, C/T), in the promoter or enhancer regions of target genes and activates transcription via the transactivation domain of p65 (155, 157).

Since NFKBIAhas a number of B binding sites in its promoter region, activated NF-B, by binding to these B sites increases the expression of NFKBIA (154)This newly synthesizedNFKBIA translocates into the nucleus and binds to NF-B, resulting in the export of

NF-B from the nucleus into the cytoplasm by means of a nuclear export sequence (NES) present on NFKBIA. This process is important to prevent prolonged activation of NF-B (160), and is further demonstrated in NFKBIA-/- mice, where lack of NFKBIA results in enhanced activation of NF-B and a severe inflammatory phenotype (161). Increased activity of NF-B has been closely associated with augmented transcription of numerous inflammatory genes (145, 152,

162). These include: pro-inflammatory cytokines such as IL1B, TNF, IL6 and CSF2; chemokines such as IL8, CCL2, CCL3, CCL5 and CCL11; inflammatory enzymes such as inducible nitric oxide synthase (iNOS) and cyclooxygenase (COX)-2 (PTGS2); adhesion molecules such as E- selectin, vascular cell adhesion molecule (VCAM)-1 and ICAM-1 (152).

23 Stimulus e.g. IL1B or TNF

IKK A IKK IKK2

B P P

Ub P P Ub Ub C

26S proteosome-mediated CYTOPLASM degradation of IB D NUCLEUS

B TATA GENE

Figure 1.2 Schematic representation of the classical, or canonical, NF-B signalling pathway. (A) Under resting conditions, NF-B, shown here as a heterodimer of p50 and p65, is sequestered in the cytoplasm by the inhibitory protein IBα (NFKBIA). (B) Upon stimulation of the cell, NFKBIA is phosphorylated by an upstream kinase, IKK2 at S32 and S36. (C) Phosphorylated NFKBIA undergoes ubiquitination and 26S proteosome-mediated degradation. (D) Degradation of NFKBIA allows nuclear translocation of NF-B heterodimer, p50 and p65, which may bind to B response elements in the promoter regions of target genes (154, 157).

Thus, NF-B is a major player in almost every aspect of the inflammatory response. In the context of asthma, increased activity of NF-B has been observed in airway epithelial cells, submucosal cells and sputum macrophages (163-165). In addition, peripheral blood mononuclear cells

(PBMCs) from severe uncontrolled asthma patients show increased expression of IKK2 and phosphorylation of NFKBIA and p65 (166). Equally, epithelial cells from asthmatic patients also

24 exhibit enhanced expression and nuclear localisation of p65 (164). Moreover, increased activity of

NF-B is also implicated in asthma exacerbations (see section 1.1.3 for further details) (167, 168).

Overall, these studies indicate that NF-B not only plays an important role in general inflammation, but may also significantly contribute to the inflammatory processes associated with asthma pathogenesis.

1.2.3.2 IRF1

The type I IFN, IFN- and -are a family of cytokines, induced upon viral infections, and are widely involved in antiviral defense, immune activation and cell growth regulation (169). By binding to their homologous receptors, IFNAR1 and IFNAR2, IFN- and - up-regulate the expression of IFN-inducible transcription factors, known as IRFs, via Janus kinase (JAK)/STAT pathway (170). The IRF family consists of 9 family members (IRF1-9) that play important roles in both innate and adaptive immunity (171). All the IRF family members have a conserved N- terminal DNA binding domain (DBD) and a C-terminal regulatory domain (172, 173). Due to the similarity of the DBD of all the IRFs, most members of the IRF family recognize same DNA binding sequence, 5’-GAAANNGAAAG/CT/C-3’ (where N denotes any nucleotide), referred to as interferon-stimulated response element (ISRE) (174). In addition to their interaction with each other as homo/heterodimers, IRFs can also co-operatively interact with other transcription factors, including NF-B and STAT, leading to enhanced gene expression (174-176) .

The first member of the IRF family, IRF-1, was originally identified as a key regulator of type I

IFN gene induction and activity (177). IRF1 mRNA is expressed in a variety of cell types, and its expression is mainly inducible by virus infection, dsRNA (poly (rl) : poly (rc)), ΙFNs (--- or cytokines, such as TNF, IL6 and IL1B (178-182). The induction of IRF1 is mainly depends

25 upon the activation of transcription factors, such as STAT1, STAT2 and NF-B (181, 183). In addition, the transcriptional activity of IRF1 is negatively regulated by its closely related IRF family member, IRF2 (170, 184). Conversely, IRF1, by enhancing the expression of STAT1 via a positive feedback mechanism, up-regulates its own expression (183, 185, 186). Even though IRF1 is a weak transcriptional activator, it induces the expression of a number of genes that are important for antiviral response/antiviral defense, apoptosis, NK cell/T cell development, macrophage function, and regulation of TH1-TH2 responses. These include: IL12p40, iNOS, CXCL10, PTGS2,

IL15, caspase 1, caspase 7, IFN-/guanylate-binding protein and protein kinase R (170, 171,

184, 187, 188).

Since IFN- induces IRF1 and plays an important role in the regulation of TH1 responses (189,

190), IRF1 could also be implicated in the regulation of TH1/TH2 response genes. In this regard, T

-/- cells from IRF1 mice, due to the decreased expression of the TH1 cytokine, IL12, and enhanced production of TH2 cytokines, failed to mount TH1 responses (191, 192). Moreover, IL1B and ΙFN-

γ-induced IRF1 also inhibits the TH2 cytokine, IL4, which further results in the attenuation of TH2 responses (193, 194). Similarly, IRF1 is also involved in Toll-like receptor (TLR)-mediated anti- viral effects (195). In addition, IRF1 plays an important role in the regulation of iNOS (196-198). iNOS catalyzes the production of nitric oxide, which, by inhibiting HRV replication, attenuates

HRV-induced production of a number of cytokines and chemokines in epithelial cells (199-202).

Conversely, since IRF1 is centrally involved in innate immunity, host defense and manipulation of cytokine responses, it may also influence the progression of asthma, where the innate immune response plays a role in the pathology (170, 192, 203, 204). In this context, genetic polymorphisms of IRF1, by inhibiting the binding of transcription factors, NF-B and early growth response

26 (EGR) 1 to IRF1 promoter, altered IRF1 mRNA and IFN- protein expression (205). Attenuated

IRF1 expression and activity was further associated with the regulation of total/specific IgE levels and the development of atopy or allergic diseases (205-208), including asthma in various ethnic groups (204, 206). However, contrary to this, enhanced IRF1 expression has been shown to reduce glucocorticoids responsiveness (209, 210) (see section 1.3.5.1 for further details). Overall, these observations suggest that IRF1 may play a critical role in the pathogenesis of asthma.

1.2.4 Post-transcriptional control of gene expression

In addition to transcriptional control, gene regulation is also controlled by post-transcriptional mechanisms. These include, mRNA stability, translation of mRNA to protein and post- translational modifications. Since these processes can influence gene regulation independently of transcriptional control, it is important to understand the role of these processes in gene expression.

However, in the context of this thesis, only the role of mRNA stability in gene regulation will be discussed below.

1.2.4.1 mRNA stability

Following the mRNA synthesis, capping and polyadenylation protects mature mRNA from being degraded by exonucleases (211). Thus, the mRNA that is spared from degradation gets exported from the nucleus into the cytoplasm, and is further translated into protein (131). The amount of mRNA templates available for translation is mainly dependent on the stability and degradation of mRNA (212). Decapping enzymes and deadenylases remove 5’-cap and 3’-poly(A) tail respectively and makes the mRNA susceptible to degradation by exonucleases (211). Moreover, sequence elements present in the 3’untranslated region (UTR) of many mRNAs can confer instability (213). The most common among these sequences is the adenylate uridylate (AU)-rich

27 sequence, also called as AU-rich element (ARE), consisting of a combination of multiple AUUUA pentamers, UUAUUUAUU nonamers or other clusters of AU pentamers and nonamers (214, 215).

Even though AREs can be found in 5’ UTR, the stabilization of mRNA is predominantly mediated by AREs present in 3’ UTR (216). The AUUUA sequences, located in the 3’ UTR of a number of inflammatory cytokines, chemokines and enzymes, including TNF, CSF2, IL6, IL8 and PTGS2 plays a critical role in the message stability (217, 218). AREs are a target for several RNA binding proteins, including tristetraprolin (TTP; zinc finger protein (ZFP) 36), human antigen R (HuR;

ELAVL1), adenine-uridine-rich element RNA-binding factor-1 (AUF1; HNRNPD) and K- homology domain splicing regulatory protein (KHSRP) (219-223). These proteins, by binding to

AREs, promote deadenylation and subsequent degradation of ARE containing inflammatory transcripts, leading to translational repression (224). For example, ZFP36 binds and destabilizes

TNF mRNA transcript, leading to translational inhibition of TNF (214). In addition, post- translational modifications, such as phosphorylation/dephosphorylation of ARE-binding proteins have also been suggested to regulate the stability of ARE containing inflammatory mRNAs. In this regard, MK2, a downstream substrate of p38 MAPK, by phosphorylating ZFP36, stabilizes inflammatory mRNA transcripts (212, 225-227). Furthermore, the destabilizing activity of AUF1 is also regulated by p38 MAPK-mediated phosphorylation (228, 229). Thus, the stability of inflammatory transcript can be controlled by a balance between stabilising and destabilising factors (212).

28 1.3 Glucocorticoids

1.3.1 Endogenous glucocorticoids

Steroid hormones are synthesized endogenously and are required for maintaining the homeostasis of the human body (230). Different classes of steroid hormones have been identified and these include; glucocorticoids, mineralocorticoids and androgens (230-232). In vivo, these steroids are synthesized from cholesterol by a stepwise series of reactions within the adrenal cortex (231, 233).

The first step of steroid synthesis involves conversion of cholesterol to pregnenolone by the enzyme cytochrome p450 (233). After which, the three groups of steroids are synthesized via distinct pathways (231, 233). Particularly for glucocorticoids, the next step of synthesis pathway involves conversion of pregnenolone to 17 OH-progesterone, which is then metabolized to 11 deoxycortisol (231). Finally, an active cortisol, the major endogenous glucocorticoid hormone, is synthesized via the metabolism of 11 deoxycortisol (231, 233, 234). The synthesis and secretion of the cortisol from the adrenal gland is tightly controlled by the HPA axis (234, 235). A number of stimuli, including infection, inflammation, pain and mental stress, activate HPA axis, leading to the release of corticotropin releasing hormone (CRH) (72). Acting on the anterior pituitary,

CRH releases adrenocorticotropic hormone (ACTH), which in turn stimulates the adrenal cortex to release cortisol (72). Once in the blood, cortisol increases the blood glucose level via the stimulation of gluconeogenesis, and also increases the metabolism of fat, protein and carbohydrates (72). In addition, cortisol also suppresses the immune system and plays an essential role in the resolution of inflammation (72). Thus, cortisol is important for maintaining and restoring the homeostasis following exposure to different stressful stimuli (236, 237). Conversely, glucocorticoids can inhibit the synthesis of CRH and ACTH (72, 236). Thus, by controlling the

29 release of cortisol from the adrenal gland, glucocorticoids participate in a negative feedback loop that regulates the level of cortisol in the blood (72, 236).

1.3.2 Molecular mechanisms of glucocorticoid action

The cellular effects of glucocorticoids are mediated by the glucocorticoid receptor, NR3C1

(nuclear receptor (NR) superfamily 3, group C, member 1), a ligand-activated transcription factor, belonging to the steroid family (238). Alternative mRNA splicing of human

NR3C1 gene generates two isoforms of NR3C1, termed  and 2. NR3C1 is a ‘classical’ ligand-activator receptor, whereas NR3C1a dominant negative inhibitor of NR3C1, is defective in steroid binding and does not bind to any known steroid agonist (240, 241). Similar to other members of the NR family, NR3C1 comprises of N-terminal transactivation domain, termed as activation function (AF) 1, a conserved DBD, and a C-terminal domain that is important for ligand binding (238, 242). The well conserved DBD of NR3C1 contains two zinc finger structural motifs and mediates DNA binding (242, 243). In addition, the DBD also comprises of the nuclear localisation sequence (NLS) and a linker region. The NLS is involved in nuclear translocation and dimerization of NR3C1 (72, 237, 243). The ligand binding domain of NR3C1 mediates the binding of glucocorticoids to NR3C1, and thus, is important in the ligand-induced activation of NR3C1

(237, 243). Furthermore, NR3C1 also contains an additional NLS2 and a second transactivation domain termed as AF2 (237, 243). Both, the constitutively active AF1 and the ligand dependent

AF2, due to their ability to interact with general transcription machinery, coactivators and chromatin remodelling complexes, regulates the transcriptional activity of NR3C1 (237, 243).

In the absence of ligand, NR3C1 is predominantly present as an inactive complex in the cytoplasm with chaperones and co-chaperone molecules (244, 245). The heat shock proteins (hsp)90 and

30 hsp70 are the two most important chaperones involved in the regulation of NR3C1 activity (244).

The binding of hsp90 to ligand binding domain of NR3C1 exposes the ligand-binding site, but masks the two NLSs and confines NR3C1 into the cytoplasm (243). The hsp90 also inhibits the binding of glucocorticoids to NR3C1 and regulates the function of the receptor in the nucleus

(246). Upon entering the cytoplasm of a cell, glucocorticoids binds to the cytoplasmic NR3C1 (72,

247), which then undergoes a conformational change with the exchange of FK506 binding protein

(FKBP)51 for FKBP52 (248). Subsequently, NR3C1 translocates into the nucleus due to the recruitment of the transport protein, dynein, by FKBP52 (248). Once in the nucleus, FKBP52 and chaperone molecule, hsp90, dissociates, leaving NR3C1 free to exert effects on transcription (Fig.

1.3) (72, 247). Activated NR3C1 binds as a homodimer to glucocorticoid response elements

(GREs) (consensus sequence; 5’-GGTACAnnnTGTTCT-3’) and further activates transcription

(transactivation) (249) (Fig. 1.3). In addition, transcriptional activation by NR3C1 requires recruitment of a number of coactivators, including the steroid receptor coactivator (SRC)1, SRC2,

SRC3, CBP/p300, PCAF and ATP-dependent chromatin remodelling complexes, to the transcriptional machinery through their interaction with AF1 and AF2 (243, 250). Binding of

NR3C1 to GREs up-regulates the transcription of GRE-dependent metabolic genes, such as tyrosine aminotransferase and phosphoenolpyruvate carboxykinase, involved in gluconeogenesis

(72). Moreover, a number of genes, including lipocortin I, secretory leukocyte protease inhibitor and the decoy IL1 type II receptor, are also up-regulated by glucocorticoids and could play significant roles in the resolution of inflammation (251, 252). However, because these proteins were induced very slowly, over 24-28 h, they cannot explain the rapid anti-inflammatory effects produced by glucocorticoids (72). Hence, up-regulation of gene transcription, or transactivation,

31 was generally believed to be associated with many of the metabolic side-effects of glucocorticoids

(72).

Following on from the characterisation of the classical GRE, very few genes, such as pro- opiomelanocortin and ACTH, were described as having a negative GRE (nGRE) sites (72). These sites were lacking some of the conserved nucleotides required for the transcriptional activation found at the classical GRE sites (72, 253, 254). Thus, initially, nGRE sites were suggested as a possible mechanism for the anti-inflammatory, glucocorticoid-dependent repression of gene transcription (62). This concept of nGRE-mediated repression is further supported by Surjit et al. suggesting that glucocorticoids can induce repression of inflammatory genes via the binding of activated NR3C1 to nGREs (255). However, this is unlikely as pro-inflammatory genes, such as

CSF2, PTGS2 and IL8, do not appear to contain nGREs in their promoters (72, 256).

1.3.3 Transrepression by glucocorticoids

An alternate mechanism that has been widely proposed for the anti-inflammatory effects of glucocorticoids is the tethering nGRE (62, 253, 254, 257). In this case, NR3C1 interacts with other transcription factors, such as NF-B and AP-1, to prevent inflammatory gene transcription

(transrepression) (Fig. 1.3) (254). Interaction of NR3C1 with these transcription factors directly represses gene transcription without involving the binding of NR3C1 to DNA (72).

Transrepression by NR3C1 is also believed to inhibit inflammatory gene transcription via the effects on DNA chromatin structure known as chromatin remodelling (see section 1.2.2 for further details) (62, 247, 254). In this context, activated NR3C1 is thought to reduce or reverse histone acetylation by the recruitment of histone deacetylases (HDACs) (254). For example, recruitment of HDAC2 by NR3C1 to NF-B leads to deacetylation of histones and subsequent inhibition of

32 Glucocorticoid

CYTOPLASM hsp90 FKBP52 + dynein

hsp90

NR3C1 FKBP51

hsp70 FKBP51 hsp70

hsp90

hsp90 dynein

NR3C1 FKBP52

NUCLEUS

NR3C1 hsp90

FKBP52 hsp90

NR3C1 NR3C1 NR3C1 TF TF

GRE TATA GENE TF-RE TATA GENE TRANSACTIVATION TRANSREPRESSION (simple GRE) (tethering nGRE) Figure 1.3. Transcriptional control by glucocorticoid receptor signalling. In the absence of glucocorticoid, glucocorticoid receptor (NR3C1) is held in the cytoplasm by chaperone molecules heat shock protein (hsp) 70, a dimer of hsp90, and FK506-binding protein (FKBP) 51. Binding of glucocorticoid to NR3C1 initiates exchange of FKBP51 for FKBP52, dissociation of hsp70 and the binding of transport protein, dynein. Subsequently, the NR3C1 complex translocates to the nuclear membrane and into the nucleus. In the nucleus, hsp 90, FKBP52 and dynein dissociate from NR3C1 to allow dimerization of NR3C1 on simple glucocorticoid response element (GRE) to initiate transcriptional activation of target genes (transactivation). Alternatively, NR3C1 may also bind to inflammatory transcription factors, such as NF-B and AP-1, and inhibit gene transcription (transrepression) (258, 259).

NF-B-mediated transcriptional responses (260, 261). Additionally, it has also been reported that

NR3C1, by inhibiting the P-TEFb-mediated phosphorylation of Pol II C-terminal domain or via the recruitment of glucocorticoid receptor interacting protein (GRIP1), represses gene transcription

33 (262, 263). These data, therefore suggest that tethering nGREs could play an important role in mediating the anti-inflammatory effects of glucocorticoids.

1.3.4 Transactivation by glucocorticoids

Accumulating evidence suggests that the up-regulation of anti-inflammatory genes by glucocorticoids can also be regarded as an important mechanism in the repression of inflammatory gene expression by glucocorticoids (114, 254, 264, 265). In this context, IL1B induced expression of IL8 and PTGS2, in A549 cells, is NF-B dependent (266, 267), and is totally repressed by dexamethasone (268). Classically, this would be explained by stating that dexamethasone represses IL8 and PTGS2 by transrepression of NF-B. However, analysis of NF-B activity indicates otherwise. In NF-B-dependent luciferase reporter cells, there was only 30-40% repression of NF-B activity by dexamethasone (268, 269). Furthermore, nuclear run-on analysis also showed a similar degree of repression for IL8 and PTGS2 by dexamethasone (268, 270).

These data therefore suggest a key role for post-transcriptional mechanism(s) in the repression of

IL8 and PTGS2. In addition to this, the repression of IL8 and PTGS2 by dexamethasone is prevented by the inhibition of transcription and translation (271, 272). This clearly suggests a role for new gene synthesis, i.e. transactivation in the repression of pro-inflammatory genes, such as

IL8 and PTGS2. Likewise, King et.al., showed that >50% of the IL1B-induced acute phase inflammatory mRNAs (peak of mRNA expression at 1 or 2 h following IL1B treatment) were repressed by dexamethasone in a manner that requires protein synthesis (273). This also suggests that transactivation by NR3C1 plays a major role in the repressive effects of dexamethasone.

To further understand the possible anti-inflammatory roles of transactivation, microarray analysis was carried out in A549 cells (271). This analysis revealed that hundreds of genes, including

34 DUSP1, ZFP36, NFKBIA, glucocorticoid-induced leucine zipper (GILZ; TSC22D3) and regulator of G-protein signaling 2 (RGS2), many with potential anti-inflammatory/bronchoprotective effects, were up-regulated by dexamethasone (271). In this context, TSC22D3 represses NF-B and AP-1-dependent transcription, whereas NFKBIA is an endogenous inhibitor of NF-B (264,

274). In addition, up-regulation of downstream of tyrosine kinase (Dok)-1, Src-like adaptor protein

(SLAP) and dexamethasone-induced Ras-related protein (Dexras)1 by glucocorticoids inhibit the activation of signalling transduction cascade through a variety of different mechanisms (275-277).

Moreover, glucocorticoid-induced expression of Clara cell secretory 10-kDa protein and thymosin

4 sulfoxide plays a protective role in allergic inflammation and neutrophilic responses respectively (278, 279). Similarly, glucocorticoid-induced RGS2, a GTPase-activating protein that attenuates G(q) signaling, is bronchoprotective (280). Thus, multiple glucocorticoid-inducible genes can regulate various aspects of the regulation of inflammatory gene expression and may therefore be expected to play redundant roles (254) (Refer to chapter 6 (section 6.3) for more details about the possible redundancy between glucocorticoid-inducible genes). In addition, glucocorticoid-induced DUSP1 and ZFP36 also play an essential role in mediating the anti- inflammatory actions of glucocorticoids and will be further discussed below.

1.3.4.1 DUSP1

DUSP1 is a nuclear phosphatase with a molecular weight of 40 kDa, induced by a number of stimuli, including pro-inflammatory cytokines, such as TNF and IL1B, LPS, glucocorticoids, and

β2-adrenoceptor agonists, such as salmeterol (114, 281-285). Inflammatory stimuli-induced

DUSP1 expression is largely MAPK dependent, and specifically, a role of ERK MAPK in the expression of DUSP1 has been demonstrated (111, 286). In addition, post-transcriptional

35 processing and post-translational modifications, including mRNA stability, phosphorylation, ubiquitination, degradation and acetylation play a crucial role in the regulation of DUSP1 activity

(287-289). In this regard, ERK-mediated phosphorylation of DUSP1 at S359 and S364 residues enhances DUSP1 stability, whereas phosphorylation at S296 and S323 residues targets DUSP1 for proteasomal degradation (287, 289, 290). In contrast, glucocorticoids, by inhibiting ERK, attenuates the proteasomal degradation of DUSP1, leading to sustained expression of DUSP1

(291). In addition, acetylation of DUSP1 does not affect protein stability or the phosphatase activity, but enhances the interaction of DUSP1 with p38 MAPK, leading to enhanced dephosphorylation of p38 MAPK and subsequently reduced expression of inflammatory genes

(288, 292). Similarly, the stability of DUSP1 mRNA in airway smooth muscle cells is enhanced by TNF in a p38 MAPK-dependent manner (293). Since DUSP1 expression is enhanced in a

MAPK-dependent manner, and enhanced expression of DUSP1 is involved in a negative regulation of MAPK signalling, DUSP1 forms a negative feedback regulatory loop to limit MAPK activity and the expression of inflammatory mediators (254, 294-297).

Inflammatory gene regulation by MAPKs often involves transcriptional and/or post-transcriptional mechanisms. In this regard, the p38 MAPK pathway is known to regulate inflammatory gene expression at the post-transcriptional level (227, 298). Thus, DUSP1-mediated inhibition of p38 is believed to predominantly contribute to post-transcriptional regulation of inflammatory gene expression by glucocorticoids (254). However, recent data also suggest that p38 also increases the transcriptional activity and/or the expression of inflammatory transcription factors including;

ATF-1, ATF-2, AP-1 and NF-B (254, 299-301). Thus, enhanced expression of DUSP1 by glucocorticoids could also regulate the transcriptional repression of inflammatory gene expression

36 by glucocorticoids (254). Similarly, DUSP1 attenuates JNK-mediated transcriptional activity of

AP-1, indicating a further role for DUSP1 in transcriptional inhibition (302). In addition, by inhibiting the synthesis of proinflammatory cytokines, DUSP1 limits the inflammatory responses in vivo (282;304;305). In this regard, DUSP1 knock-out mice exhibit enhanced MAPK activity with exaggerated inflammatory responses in response to LPS or zymosan (281, 303, 304).

However, the repressive effect of glucocorticoids in DUSP1-/- mice was gene specific. For example, while glucocorticoid-induced repression of CSF2 and IL1A was strongly attenuated, the repression of TNF, IL1B and PTGS2 was only partially attenuated (281). These data, therefore imply that additional, DUSP1-independent mechanism of repression by dexamethasone must exist for the inhibition of inflammatory genes (281, 295).

1.3.4.2 ZFP36

ZFP36 is an mRNA destabilizing protein, belonging to the family of tandem CCCH zinc finger proteins, and its expression is enhanced in T cells, macrophages, fibroblasts and pulmonary A549 cells by a number of stimuli, including IL1B, LPS, phorbol esters, growth factors and β2- adrenoceptor agonists, such as salbutamol (305-308). Pro-inflammatory stimuli-induced ZFP36 protein expression is highly dependent on MAPK activation (308, 309). ZFP36 protein appears initially as a ~40 kDa protein, which becomes phosphorylated and migrates at ~45-kDa on SDS-

PAGE (308, 310). Phosphorylation of ZFP36 at S52 and S173 residues is mediated by a p38- dependent kinase, MK2, and is suggested to enhance ZFP36 stability (310). This stabilisation of

ZFP36 is mediated via the binding of the adaptor protein, 14-3-3, to ZFP36 (310), which in turn inhibits the destabilizing activity of ZFP36 (311). In contrast, dephosphorylation of ZFP36 by

PP2A increases the activity of ZFP36 (312). Thus, a regulatory loop exists whereby an initial

37 MK2-mediated phosphorylation of ZFP36 in the presence of inflammatory stimuli results in

ZFP36 with low activity to allow the up-regulation of inflammatory transcripts by MAPKs.

However, the subsequent dephosphorylation of ZFP36 by PP2A activates mRNA-bound ZFP36, leading to a rapid degradation of target mRNA, thus preventing prolonged or exaggerated inflammatory responses (226).

Given the ability of ZFP36 to reduce the expression of ARE-containing mRNAs (see section

1.2.4.1 for more details), it is often suggested that ZFP36 is a negative feed-forward regulator of inflammatory gene expression (305). In this regard, ZFP36 is an established negative regulator of

ARE-containing mRNAs, including TNF, PTGS2, CSF2, IL6 and IL8, and mice lacking ZFP36 develop severe and chronic inflammation (215, 222, 313). The mRNA destabilizing effect of

ZFP36 is mainly attributed to its interaction with the mRNA decay machinery, such as decapping enzymes, the exonuclease, XRN1, and the subunits of the exosome (314, 315). Conversely, the deadenylation of poly(A) tail by ZFP36 is mainly due to the activation of the deadenylase, poly(A)- specific ribonuclease (314-316).

The expression of ZFP36 is also modestly up-regulated by glucocorticoids in primary human airway epithelial cells, pulmonary A549 cells, bronchial BEAS-2B epithelial cells and in the airways following glucocorticoid inhalation (Leigh et. al. unpublished data) (306, 308, 317).

Furthermore, the role of ZFP36 in glucocorticoid-induced repression of inflammatory gene expression is also indicated (295, 317). In this regard, glucocorticoid-induced repression of chemokine gene expression was impaired in ZFP36 knockout mice (317). Moreover, siRNA- mediated knock-down of ZFP36 attenuated dexamethasone-induced repression of TNF and IL8

(306, 318). Conversely, LPS and IL1B-induced ZFP36 expression is partially repressed by

38 dexamethasone (308, 319). In this context, recent data point to the possibility that the dexamethasone-mediated repression of ZFP36 could be due to the enhanced expression of DUSP1 in the presence of dexamethasone, which, by inhibiting p38 MAPK, reduces the expression of

ZFP36 (320). However, this hypothesis warrants further investigation and is further tested in this thesis.

1.3.5 Insensitivity/resistance to glucocorticoids in asthma

Although glucocorticoids are widely used for treating asthma, symptoms related to severe asthma are difficult to control due to poor sensitivity or resistance to the anti-inflammatory effects of glucocorticoids (71). Glucocorticoid resistance may refer to a decrease in the maximum response to glucocorticoids, a rightward shift in the dose response curve or potentially both of these effects

(321, 322). Clinically, glucocorticoid resistance is defined as the presence of unresolved airway obstruction or no improvement in lung functions, such as forced expiratory volume (FEV1) or PEF, after two weeks of treatment with high dose oral glucocorticoids (62, 322). Overall, resistance to glucocorticoids complicates the clinical management of asthma and results in a significant financial burden (71, 247). Further understanding of the molecular mechanisms of glucocorticoid resistance may help in the development of an effective anti-inflammatory therapy for the treatment of severe asthma (322).

1.3.5.1 Mechanisms of reduced glucocorticoid sensitivity

A number of mechanisms may contribute towards the decreased responsiveness of severe asthmatics to glucocorticoids (247, 321, 322). In this context, TNF plays an important role in asthma pathogenesis and enhanced level of TNF is often associated with reduced sensitivity to glucocorticoids (323, 324). The expression of TNF is increased in BAL and bronchial biopsy

39 specimens from severe asthma patients, and was further linked with glucocorticoid resistance

(325). In addition, TNF reduces glucocorticoid responsiveness in monocytes (326) and activates pathways that are involved in chronic airway remodeling and sub-epithelial fibrosis (327).

Moreover, enhanced activation of ERK and p38 MAPK pathway by TNF enhances the expression of IL8, a neutrophil chemoattractant (328). Activation of ERK MAPK in airway smooth muscle cells is insensitive to the repressive actions of glucocorticoids, and is further involved in the recruitment of neutrophils contributing to airway inflammation (328). Additionally, cytokines associated with TH1 immunity may also contribute to the pathogenesis of severe glucocorticoid resistant asthma (329). For example, IFN-/TLR4-MyD88-dependent AHR is resistance to glucocorticoid therapy (330). Similarly, treatment of murine macrophages with IFN-inhibits nuclear translocation of NR3C1 through the activation of MyD88 pathway (331). Furthermore, increased exposure to cytokines and activation of MAPK have also been found to induce glucocorticoid insensitivity through attenuating NR3C1 ligand binding (322). For example, the

TH2 cytokines, IL2, IL4 and IL13, have been found to decrease the binding of glucocorticoids to

NR3C1 (321). In addition, cytokines, such as TNF and TGF-, by decreasing the expression of

NR3C1, reduces glucocorticoid responses (326, 332). Equally, p38 and JNK MAPK-mediated phosphorylation of NR3C1 reduces the translocation of NR3C1 into the nucleus and suggested to contribute to glucocorticoid insensitivity (333). In this respect, cytokines, such as IL2 and IL4, by activating p38 MAPK, decreases nuclear translocation of NR3C1 (247). In addition, enhanced activity of inflammatory transcription factors, such as NF-B, STAT5, IRF1 and AP-1, have all been implicated in reduced response to glucocorticoids (322). STAT5 has been found to bind

NR3C1 and further associated with defective nuclear translocation and DNA binding of NR3C1

(334). AP1 is thought to interact with monomeric NR3C1 and thereby inhibiting the interaction of

40 NR3C1 with GRE or other transcription factors (335). The expression of NF-B was enhanced in

PBMCs from asthmatics and has been shown to correlate with reduced glucocorticoid responsiveness (336). IFNs and TNF-mediated activation of IRF1 may also contribute to reduce response to glucocorticoids in airway smooth muscle cells through sequestration of the transcriptional co-activator of NR3C1, GRIP1 (209, 210). In this context, depleting the amount of

GRIP1 available to NR3C1 in the presence of high level of IRF1 would decrease the transcriptional function of NR3C1 (209, 210).

The detailed nature of the asthmatic phenotype may also play a crucial role in determining the responsiveness of different asthmatics to glucocorticoid therapy. Thus, even though glucocorticoids are highly effective in reducing eosinophilic inflammation, as seen in mild to moderate asthmatics, severe asthmatics, with neutrophilic inflammation, respond poorly to glucocorticoids (321, 337). Furthermore, enhanced activity of phosphoinositide 3-kinase (PI3K), especially in asthmatics who smoke, by attenuating HDAC activity, may contribute to reduced glucocorticoid sensitivity (338). In this context, smokers also show a marked reduction in HDAC2 activity (339, 340). Since, HDAC2 is important for the repression of NF-B by glucocorticoids

(260, 261), attenuation of HDAC2 activity may further result in reduced responses to glucocorticoids. However, contrary to this, recent work by Wang et al. demonstrated that budesonide-mediated inhibition of cytokine release in patients with severe smoking history was independent of HDAC2 activity, and thus, inhibition of HDAC2 activity may not always corresponds to reduced response to glucocorticoids (209, 341). Further, increased expression of

NR3C1β, a dominant negative inhibitor of NR3C1, by interfering with DNA binding ability of

NR3C1, reduces glucocorticoid responsiveness (342). NR3C1β by binding to GRE may out-

41 compete NR3C1 binding to GRE (322, 343). NR3C1β has also been found to form heterodimers with NR3C1, which may inhibit NR3C1 interaction with GRE (241). However, the data supporting the role of NR3C1in reduced glucocorticoid response are not clear and currently inconclusive. Overall, insensitivity to glucocorticoids may involves a number of cellular and molecular mechanisms, and an improved understanding of these mechanisms may help to identify new targets for therapeutic interventions.

1.4 General hypothesis and research aims

1.4.1 General hypothesis

The overriding hypothesis for this thesis is that glucocorticoids induce the expression of genes, which then play critical roles in the repression of inflammatory gene expression. In addition, this thesis also seeks to test the hypothesis that some glucocorticoid-induced genes may also play an opposing role by promoting resistance/insensitivity to the repressive effects of glucocorticoids.

1.4.2 Research aims:

The overall aims of this thesis are: (I) to investigate the roles of glucocorticoid-induced DUSP1 in the repression of inflammatory gene expression by glucocorticoids; (II) to examine the feed- forward inhibitory role of ZFP36 on the regulation of ARE-containing inflammatory transcripts in

DUSP1 inhibited cells and implication of such effector mechanisms on glucocorticoid-induced repression of inflammatory genes; and (III) to study glucocorticoid-induced DUSP1 as a mechanism of glucocorticoid insensitivity.

42 Chapter 2 : Material and Methods 2.1 Materials

AbD Serotec, Raleigh, NS, USA: Primary antibodies (see Table A1)

Abcam, Toronto, Canada: Primary antibodies (see Table A1)

American Type Culture Collection, Rockville, MD, USA: A549 cells

Applied Biosystems Inc, Foster City, CA, USA: MicroAmp Optical 96 well reaction plate

Biotium Inc, Hayward, CA, USA: Luciferase assay buffer

Calbiochem, Ontario, Canada: SB203580, UO126, JNK inhibitor 8

Cell Signalling, Danvers, MA, USA: Primary antibodies (see Table A1)

Dako, Mississauga, ON, Canada: Goat anti-rabbit, goat anti-mouse and rabbit anti-goat immunoglobulins

Diagenode, Denville, USA: Bioruptor/sonicator

Fisher Scientific, Nepean, ON, Canada: Methanol, ethanol, sulphuric acid (H2SO4), glacial acetic acid, hydrochloric acid (HCI), sterile tissue culture reagents

Invitrogen, Burlington, ON, Canada: Dulbecco's Modified Eagle's Medium (DMEM),

Dulbecco's Modified Eagle's Medium (DMEM) F-12, fetal calf serum (FCS), NuPage western blotting system, Lipofectamine RNAiMAX, Lipofectamine 2000, trypsin ethylenediaminetetraacetic acid (EDTA), penicillin-streptomycin, SYBR greenER mastermix,

Fast SYBR green master mix, protein G magnetic Dynabeads, proteinase K

43 Thermo Scientific - Pierce Protein Research Products, Rockford, IL, USA: ECL western blotting substrate, 16% formaldehyde, protease inhibitor cocktail

Promega, Madison, WI, USA: Reporter lysis buffer, magnetic separation stand (twelve-position) for ChIP

Qiagen Ltd, Mississauga, ON, Canada: QlAshredder, RNeasy Mini Kit, DNase set including

RDD buffer, siRNAs (see Table A2)

R&D Systems, Minneapolis, MN, USA: Human recombinant IL1B, human recombinant TNF, duoSet CSF2, CXCL1, CXCL10, IL6, IL8 and TNF ELISA kits

Roche Diagnostics, Laval, QC, Canada: Complete protease inhibitor tablets for western blot analysis

Santa Cruz Biotechnology Inc., Santa Cruz, CA, USA: Primary antibodies (see Table A3)

Sigma-Aldrich Company, Oakville, ON, Canada: TRIZMA-base, L-glutamine, phosphate- buffered saline (PBS), Hanks balanced salt solution (HBSS), phenylmethylsulphonyl fluoride

(PMSF), sodium orthovanadate (Na3VO4), sodium fluoride (NaF), sodium pyrophosphate

(Na4P207), bovine serum albumin (BSA), orange G, ethidium bromide, dexamethasone, sodium bicarbonate (NaHC03), polyoxyethylene-sorbitan (Tween 20), glycine, Ponceau S, bromophenol blue, β-mercaptoethanol, select agar, acrylamide-bis, TEMED, 1,4-Piperazinediethanesulfonic acid, Piperazine-1,4-bis(2-ethanesulfonic acid) (PIPES), Corning Spin-X® UF 500 Concentrator

5000 columns, sodium chloride (NaCl), sodium deoxycholate, sodium dodecyl sulphate (SDS),

Nonidet P-40 (NP-40)

44 2.2. Methods

2.2.1 Cell Culture methods

2.2.1.1 A549 cell culture

A549 cells are adenocarcinoma human alveolar basal epithelial cells derived via culturing of cancerous lung tissue in the explanted tumor of a 58-year-old Caucasian male (344). The cells were grown in 175 cm2 culture flasks in a media containing Dulbecco’s modified Eagle’s medium

(DMEM) supplemented with 10% v/v foetal calf serum (FCS) and 2 mM L-glutamine. Cells were

o incubated at 37 C, 5 % CO2/95% air and passaged when 90-95% confluent. For experiments, cells were cultured either in 6, 12 or 24 well plates and incubated with serum-free medium, to arrest the cell growth, overnight prior to experiments.

2.2.1.2 Primary human bronchial epithelial (HBE) cell culture

Primary human bronchial epithelial (HBE) cells were collected and cultured in the laboratory of

Dr. David Proud (University of Calgary, Canada). HBE cells isolated from non-transplanted normal human lungs obtained using the tissue retrieval service at the International Institute for the

Advancement of Medicine, Edison, NJ, USA, were cultured in bronchial epithelial cell growth medium (BEGM) supplemented with 5% FCS for the first 72 h of culture at 37°C in 5% CO2/95% air. After 72 h, cells were washed with Hank’s balanced salt solution (HBSS) and fed with BEGM for further 14 days. The medium was changed every two days. After ~14 days, once the cells had reached >85% confluence, they were used for experiments. For experiments, cells were cultured in 6 well plates in bronchial epithelial cell basal medium (BEBM) and incubated with BEBM, to arrest the cell growth, overnight prior to experiments.

45 2.2.1.3 Adenoviral infection

A549 cells were grown in 6 or 12 well plates to approximately 70% confluence. Cells were then infected with multiplicity of infection (MOI) of 10 of either an empty adenoviral serotype 5 (AD5) green fluorescent protein (GFP)-expressing vector (Ad5-GFP) or Ad5-DUSP1, containing a

DUSP1 cDNA expression cassette (345) for 24 h in serum containing media. After 24 h cells had reached full confluence and were then incubated in serum-free media overnight, to arrest the cell growth, prior to experiments.

2.2.1.4 siRNA-mediated gene silencing

Gene silencing was carried out by using DUSP1-, ZFP36-, IRF1- or lamin A/C (LMNA)-specific siRNAs at a final concentration of 25 nM (DUSP1 & ZFP36 siRNA) or 50 nM (IRF1 siRNA).

Concentration of control LMNA siRNA was same as that for the targeting siRNAs. Each siRNA was mixed with lipofectamineTM RNAiMAX (1 μl of 1 μg/μl) (Invitrogen) in 100 μl of serum-free

DMEM and incubated at room temperature for 30 min prior to dilution to 1 ml and addition to cells. A549 cells were grown to approximately 60-70% confluence in 12 well plates before being incubated for 24 h at 37oC in serum free DMEM containing siRNA and lipid. Cells were then treated with specific stimuli for an experiment. LMNA siRNA was used as a control siRNA and was subjected to an identical protocol to the siRNA of interest. The LMNA-targeting siRNA- mediated knockdown of LMNA was confirmed previously in our laboratory (308). See Table A1 for siRNA sequences.

46 2.2.2 Western blotting

2.2.2.1 Preparation of cell lysates

Cells were lysed in 1 × Laemlli buffer (10% v/v glycerol, 5% v/v β-mercaptoethanol, 0.625 M

SDS, 0.25 M TRIZMA-HCL pH 6.8 and 0.75 mM bromophenol blue) containing 1 × completeTM protease inhibitor mixture and phosphatase inhibitors (50 mM NaF, 2 mM Na3VO4 and 20 mM

Na4O7P2). To achieve complete cell lysis, cells were frozen overnight at -20°C. After defrosting, cells were scraped from plates on ice and transferred to a fresh 1.5 ml Eppendorf tube. The cell lysates were then heated to 100oC for 10 min and subjected to sonication for 30 min at 4oC. After sonication, the samples were again heated to 100oC for 2 min before loading into gels.

2.2.2.2 SDS polyacrylamide gel electrophoresis

Gel electrophoresis of proteins was carried out by using the Novex NuPAGE western blotting system (Invitrogen, Paisley, U.K.) or Bio-Rad Mini-PROTEAN gel electrophoresis system. Cell lysates (25 to 30 µl) and rainbow protein ladder (5 µl) were loaded on to the gel. Total proteins were size-separated at 150 V for 90 min on 10-12% polyacrylamide gel by using 1 × running buffer

(25 mM TRIZMA base, 191.8 mM glycine, 0.1% w/v SDS).

2.2.2.3 Protein transfer to nitrocellulose membranes

Proteins were then electro-transferred onto nitrocellulose membranes using 1 × transfer buffer (25 mM TRIZMA base, 191.8 mM glycine, 20 % v/v methanol) at 0.4 mA for 90 min on ice. Following the transfer of proteins onto the membranes, equal loading of protein was confirmed by staining the membranes with Ponceau S staining dye (1.3 mM Ponceau S dissolved in 5% v/v acetic acid) for 1 min at room temperature. Destaining to visualise protein bands was achieved by washing the

47 membranes with 1 × washing buffer (1 × TBS-Tween, 0.5 M TRIZMA base, 1.5 M NaCl, 0.1 % v/v Tween 20) for 5 to 10 min with agitation on an orbital shaker.

2.2.2.4 Immunodetection of proteins

After staining, the membranes were blocked in 1 × TBS-Tween containing 5% skimmed milk for

1 h at room temperature with continuous agitation on an orbital shaker. The membranes were then washed for 5 min with 1 × washing buffer and incubated with appropriate polyclonal primary antibodies according to the manufacturer’s guidelines, between few hours to overnight, either at room temperature or at 4oC in TBS-Tween plus 5% BSA or skimmed milk. The membranes were then again washed for 30 min with 1 × washing buffer with gentle agitation on an orbital shaker.

Washing buffer was changed after every 10 min. The membranes were then incubated with anti- rabbit, anti-mouse or anti-goat immunoglobulins linked to horseradish peroxidase diluted in TBS-

Tween plus 3% skimmed milk for 1 h at room temperature. Following 30 min of further washing

(3 x 10 min each wash), immune complexes were detected using enhanced chemiluminescence

(ECL) and visualised by autoradiography. The saturation of immune complexes was avoided by taking multiple exposures of x-ray films. Autoradiographic bands were quantified by densitometry using Total LabTM software (Nonlinear Dynamics, Newcastle, UK). See Table A2 for details of antibodies used.

2.2.3 RNA isolation, cDNA synthesis and SYBR Green Real Time PCR

2.2.3.1 RNA isolation

Total RNA was extracted using the RNeasy mini kit (Qiagen) according to manufacturer’s guidelines. Cells were lysed in 350 L of RLT buffer containing 1% v/v -mercaptoethanol and frozen at -80oC. Cells were thawed and scrapped from plates on ice. The lysates were then

48 transferred into a QIAshredder spin column and centrifuged at maximum speed for 2 min. The homogenized lysates were then mixed with an equal volume (350 L) of 70 % ethanol, applied to

Rneasy mini column and centrifuged at 9279 g (10000 rpm) for 30 sec. The flow through was discarded and the mini columns were then washed with 350 L of buffer RW1. DNA contamination was removed by treating the spin columns with 80 l of 1 × RDD buffer containing

18 units of DNase I and incubation at room temperature for 30 min. The columns were then washed once with 350 L RW1 and twice with 500 L buffer RPE for 30 sec at 9279 g. For the second wash with buffer RPE, the samples were centrifuged for 2 min at 9279 g. The flow through was discarded and the spin columns were further centrifuged at full speed for 1 min to remove residual

RPE. The columns were then transferred to 1.5 ml collection tube. To elute the RNA, 30 l of

RNase-free water was added directly onto silica gel membrane and incubated at room temperature for 1-2 min before centrifugation at ≥ 9279 g (10000 rpm) for 1 min. Eluted RNA was quantified by measuring the absorbance at 260 nm using a NanoDrop instrument and stored at -80oC.

2.2.3.2 cDNA synthesis cDNA synthesis was carried out using Quanta bioscience qscript cDNA synthesis kit. 0.5 g RNA was reverse transcribed into cDNA in 20 l reaction mixture containing 1 × Moloney Murine

Leukemia Virus (MMLV) reverse transcriptase, 1 mM each dNTPs [dATP, dCTP, dTTP, dGTP],

1 U RNase inhibitor and 5 ng random hexamers plus oligo (dT). Reaction cycling parameters for

Eppendorf thermocycler were 70oC for 5 min, 22oC for 5 min, 42oC for 30 min and 90oC for 4 min. The resultant cDNA was then diluted 1:4 in Rnase free water and stored at -80oC.

49 2.2.3.3 SYBR green real-time PCR

Real-time PCR (RT-PCR) was carried out using ABI 7900HT or StepOnePlus™ instruments. Each

RT-PCR reaction contained 2.5 µl of cDNA plus 5 l of SYBR greenER MasterMix or Fast SYBR green MasterMix, 2.3 l of water and 2.5 ng of each of the appropriate forward and reverse primers in a 10 l of final reaction volume. Relative cDNA concentrations were obtained from standard curves generated by a serial dilution of stimulated cDNA sample analysed at the same time as experimental samples. Amplification conditions were: 50oC for 2 min, 95oC for 10 min then 40 cycles of 95oC for 15 s and 60 or 70oC for 1 min. Primer specificity was assessed using dissociation

(melt) curve analysis: 95oC for 15 s, 60oC for 20 s followed by ramping to 95oC over 20 min. The specificity of primers was indicated by a single peak in the change of fluorescence with temperature plot. See Table A3 for primer sequences.

2.2.3.4 Analysis of unspliced nuclear RNA

Unspliced nuclear RNA, or nascent transcript, accumulates transiently in the nucleus following transcriptional activation and may be measured as a surrogate of transcription rate (346, 347).

Unspliced nuclear RNA was analyzed using SYBRgreen primers that crossed the exon 3/intron 4 junction for TNF and exon 1/intron 1 junction for IRF1. Since the primer sets detect both unspliced

RNA and genomic DNA, the signal due to the contaminating genomic DNA was assessed in each sample. Each RNA sample was subjected to reverse transcription, both in the absence and the presence of reverse transcriptase. The presence of an amplification product in the reverse transcription-negative samples was attributed to genomic DNA contamination and samples with greater than 10% genomic contamination for U6 were excluded from further analysis. RNA

50 extraction, cDNA synthesis, and SYBR green RT-PCR were carried out as described above. See

Table A3 for primer sequences.

Since unspliced GAPDH RNA was down-regulated by TNF treatment (Newton R. unpublished data), small nuclear RNA U6 was used as the housekeeping gene for nuclear RNA normalisation.

Unspliced U6 RNA expression did not change upon stimulation of cells (Newton R. unpublished data). U6 is RNA polymerase III dependent small non-coding RNA (snRNA), and a component of the spliceosome that is expressed in the nucleus (348).

2.2.4 Enzyme-linked immunosorbent assay (ELISA)

Cell supernatants were collected and the cellular debris were removed by centrifugation at 2320 g

(5000 rpm) for 2 min at 4oC. The supernatants were then transferred to fresh 1.5 ml Eppendorf tubes and stored at -20oC. ELISA for CSF2, CXCL1, IL6, IL8, TNF and CXCL10 was carried out using R&D systems DuoSet ELISA kits according to the manufacturer’s guidelines. Flat bottomed

96-well ELISA plate was coated using 100 l of capture antibody. The plate was then incubated overnight at room temperature and washed 3 times with washing buffer (PBS-0.1% v/v Tween

20). After blocking with reagent diluent (PBS containing 1 % w/v BSA) for 1 h, the plate was again washed 3 times. 100 l of cell supernatants or standards was then loaded to the plate and incubated for 2 h at room temperature. Following further 3 washes, 100 l of biotinylated detection antibody was added to the plate and incubated for another 2 h. The plate was then washed 3 times and 100 l of streptavidin horseradish peroxidise was added. After incubating for further 20 min, the plate was washed again 3 times and 100 l of H2O2/tetramethylbenzidine substrate solution was added. The colour reaction was developed by incubating the plate at 37oC in the dark. The development reaction was stopped by adding 50 l of 2N H2SO4 to each well and sample

51 absorbance was read at 450 nm on a FLUOstar optima plate reader. Protein concentrations in the samples were determined through interpolation of a linear standard curve of known protein concentrations.

For HBE cells, TNF release was determined following concentration of 900 μl of cell supernatant to less than 100 μl using Corning Spin-X® UF 500 Concentrator 5000 columns. Concentrated supernatants were adjusted to final volume of 110 l using 1% w/v BSA/PBS, and ELISA for TNF was performed using 100 l of concentrated supernatant as described above. Cell-associated TNF in A549 cells was detected by lysing the cells in 100 l of 1 × firefly luciferase assay buffer containing 1 × CompleteTM protease inhibitor mixture and phosphatase inhibitors (50 mM NaF, 2 mM Na3VO4 and 20 mM Na4P2O7). ELISA for TNF was carried out using 50 l of cell lysates diluted with 50 l of 1% w/v BSA/PBS. Standard curves were generated using the same firefly luciferase assay buffer and ELISA was carried out as described above.

2.2.5 Chromatin Immunoprecipitation (ChIP) assay

ChIP was performed to study the interaction between the transcription factor, IRF1, binding to inflammatory gene promoter sites (protein/DNA interaction) following differential cell treatment.

2.2.5.1 Cell fixation

Following desired treatment of A549 cells, grown to confluence in 100 mm dishes, the cells were fixed to cross-link and preserve protein/DNA interactions. Cross-linking was achieved by adding

500 l of fixation solution (0.75% v/v formaldehyde: 375 l 16% formaldehyde + 125 l of double distil water) directly to the culture medium and incubating for 10 min at room temperature with gentle rocking on an orbital shaker. Formaldehyde was quenched by adding 405 l of 2.5 M

52 glycine (final concentration 125 mM in total of 8 ml solution) and incubating for 5 min at room temperature with gentle rocking on an orbital shaker. The fixation solution was poured off and the cells were further washed with 5 ml of ice-cold 1 × HBBS for 5 min at room temperature with gentle rocking on an orbital shaker. HBSS was aspirated completely from the dish and 4 ml of ice- cold immunoprecipitation (IP) lysis buffer (5 mM PIPES at pH 8.0, 1 mM EDTA, 85 mM KCl,

5% v/v glycerol, 0.5% v/v NP-40) supplemented with protease inhibitor cocktail (PIC) was added.

The cells were scraped from the dish, collected in a pre-chilled 15 ml conical tube and lysed by nutating for 2 h at 4°C. The nuclei were collected by centrifugation (600 g for 5 min at 4°C) and resuspended in 2 ml of IP lysis buffer supplemented with PIC. The lysed cells were then passed through a 5 ml syringe with 30G needle to release the nuclei completely and collected in the same chilled 15 ml conical tubes. The cells were further centrifuged (600 g for 5 min at 4°C), supernatants were discarded and nuclei were resuspended in 300 l of ice-cold radioimmunoprecipitation assay (RIPA) buffer (1 × PBS, 1 mM EDTA, 150 mM NaCl, 5% v/v glycerol, 0.5% w/v sodium deoxycholate, 0.1% w/v SDS, 1% v/v NP-40) supplemented with PIC.

The nuclei were either stored at -80°C or the protocol was continued.

2.2.5.2 Chromatin shearing by sonication

Chromatin sheared to a size of 200-500 bp was used for ChIP experiments. Frozen nuclei samples were thawed on ice and sonicated with a Diagenode Bioruptor at high power in 30 sec bursts separated by 30 sec incubations in ice-cold water maintained at 4°C for a total of 30 min. Lysates were cleared by centrifugation at maximum speed for 15 min at 4°C on a table-top centrifuge. Of

300 l sheared nuclei supernatants, 270 l of supernatants were transferred to a fresh pre-chilled

1.5 ml Eppendorf tube and used for immunoprecipitation. The remaining 30 l of supernatants

53 was transferred to a separate pre-chilled 0.2 ml PCR tube, stored at -80oC and used as input DNA to check the sonication efficiency and for the validation of ChIP primers.

In order to check the sonication efficiency, 90 l of cross-link reversal solution (TE (Tris-EDTA) containing 0.7% w/v SDS, 200 g/ml proteinase K and 5 M NaCL) was added to 10 l of input

DNA and cross-links were reversed for 3 h at 55°C followed by 16 h at 65°C using an Eppendorf thermal cycler. Input DNA was then purified using ChIP DNA clean and concentrator kit (Zymo

Research) using 7 volumes of DNA binding buffer, followed by elution in 50 l elution buffer.

Cleaned input DNA was then run on 1 % agarose gel for 45 min-1 h at 80 V and sizes of sheared

DNA fragments were visualised using a gel imager.

2.2.5.3 Immunoprecipitation of sheared chromatin

After the samples had been subject to shearing by sonication, the immunoprecipitation reaction was set up. Protein G magnetic Dynabeads were used for immunoprecipitation. Beads were mixed thoroughly by vortexing for 30 s on the vortex and sufficient quantities of beads (100 l/ IP reaction) were transferred to a pre-chilled 1.5 ml Eppendorf tube. Beads were then pelleted by placing the tubes on magnetic stand. The liquid was removed by pipetting without disturbing the beads and the beads were then washed quickly two times with ice-cold 500-600 l of RIPA buffer, with resuspending the beads each time. After the final wash, the beads were resuspended in RIPA containing 1 × PIC and BSA (5 mg/ml) at original volume (i.e. if you have started with 100 μl beads, resuspend them in 100 μl). To this, 12 g of IRF1 (C-20, SC-497) antibody was added and nutated overnight at 4°C.

54 Antibody-bound beads were pelleted with magnetic stand and washed quickly two times with ice- cold 500-600 l of RIPA buffer, with resuspending the beads each time. After the final wash, the beads were resuspended in RIPA containing 1 × PIC and BSA (1 mg/ml) at original volume. 100

l of antibody-bound beads were then transferred to each tube containing 270 l of sheared chromatin DNA, thawed earlier on ice, and nutated overnight at 4°C.

2.2.5.4 Washing and cross-link reversal of ChIP DNA

Beads were collected using a magnetic stand and subjected to four washes with ice-cold RIPA buffer containing 500 mM NaCl, followed by 4 washes with ice-cold LiCl buffer (20 mM Tris at pH 8.0, 1 mM EDTA, 250 mM LiCl, 0.5% v/v NP-40 and 0.5% w/v sodium deoxycholate). After the final wash, beads were pelleted using magnetic stand and the liquid was removed from each tube. The cross-link was then reversed using Eppendorf thermal cycler. In order to reverse the cross-link, 100 l of cross-link reversal solution (TE (Tris-EDTA) containing 0.7% w/v SDS, 200

g/ml proteinase K and RIPA w/500 mM NaCL) was added to antibody-bound beads at room temperature and the samples were transferred to a fresh 0.2 ml PCR tube. The conditions for reversing the cross-link were: 3 h at 55°C followed by 16 h at 65°C.

2.2.5.5 ChIP DNA clean-up

After reversing of the cross-links, ChIP DNA was purified using ChIP DNA clean and concentrator kit (Zymo Research). Cross-linked reversed antibody-bound beads were pelleted and the supernatants were transferred to a fresh 1.5 ml Eppendorf tube. To this, 7 volumes of DNA binding buffer was added, mixed well before transferring to Zymo-Spin columns and centrifuged at ≥ 9279 g (10000 rpm) for 30 s. The flow through was discarded and the columns were then washed twice using 250 l of washing buffer by centrifugation at ≥ 9279 g (10000 rpm) for 30 s. To elute the

55 ChIP DNA, 50 l of elution buffer was added directly to the columns matrix and incubated at room temperature for 1-2 min before centrifugation at ≥ 9279 g (10000 rpm) for 30 s. Eluted ChIP DNA was then further used for PCR or stored at -20°C.

2.2.5.6 SYBR green PCR using ChIP DNA qPCR amplification and SYBR green detection method was employed to analyse the DNA obtained by ChIP. IRF1 occupancy for a given region under a specific experimental condition was measured as the difference between the CT value for the specific region relative to the geometric mean of the CT values for three negative control regions (hMYOD1, hOLIG3, hMYOG) not predicted to occupy IRF1. Amplification of diluted input DNA generated similar CT values for control and test regions primers. Assays were generally performed in biological replicate and repeated at least 4 times with qualitatively similar results. Primer sequences are shown in Table

A4. PCR conditions were same as described earlier in section 2.2.3.3.

2.2.6 3-(4, 5-dimethylthiazol-2-yl)-2, 5-diphenyltetrazolium bromide (MTT) assay

MTT assay was used to assess the viability of cells following various treatments. This assay is based on the ability of a mitochondrial dehydrogenase enzyme, present in viable/proliferating cells, succinate dehydrogenase, to cleave the tetrazolium rings of MTT. MTT is pale yellow in color and is converted to dark blue/purple formazan crystals upon cleavage by the mitochondrial dehydrogenase enzyme. The number of viable cells is directly proportional to the amount of formazan product/blue purple color in the wells (349).

Following the desired treatments of A549 cells in a 12-well plate, the supernatant was aspirated,

200 μl/well of MTT reagent (1 mg/ml in HBSS) was added and incubated at 37°C for 30 min. The

MTT reagent was aspirated from the wells and the formazan crystals were then solubilised by the

56 addition of DMSO (200 μl/well). 100 μl of each sample was then transferred into a 96 well plate and the colour was quantified by determining the optical density units (ODU) at 584 nm using a

FLUOstar optima plate reader. Percent viability of the cells was calculated by comparing to medium control using the following formula:

% viability = ODU (sample) × 100 ODU (medium control)

2.2.7 Data presentation and statistical analyses

GraphPad Prism 5 software was used for all statistical analyses. All graphical data are plotted as mean ± S.E. and normalised to either GAPDH (protein and cytoplasmic RNA) or U6 (unspliced nuclear RNA) expression as indicated. One-way ANOVA with a Bonferroni post-test was used for comparing 5 or fewer comparisons. Since the Bonferroni post-test gives high and increasingly inappropriate false negative rates (i.e. type II or β error) for greater than five comparisons, ANOVA with Newman-Keuls multiple comparison test was used for greater than five comparisons, as is recommended for greater power in hypothesis testing (Prism 5, Graphpad Software). ANOVA with a Dunnett's post-test was used for comparisons against a single control column. Two-tailed, paired Student t test was used for comparing two treatment groups. ChIP data were analyzed by

Friedman test with Dunn’s post-test. Significance between groups was assumed as follows, P <

0.05 (*), P < 0.01 (**), P < 0.001 (***).

57 Chapter 3 : Feedback control of inflammatory gene expression by the mitogen-activated protein kinase (MAPK) phosphatase, DUSP1, and regulation by dexamethasone

Data presented in this chapter have been published:

Shah S, King EM, Chandrasekhar A, Newton R. Roles for the mitogen-activated protein kinase (MAPK) phosphatase, DUSP1, in feedback control of inflammatory gene expression and repression by dexamethasone. J. Biol. Chem. 2014;289(19): 13667-13679

Copyright  Journal of Biological Chemistry.

This work was reprinted with permission based on the J. Biol. Chem. copyright permission policy.

King EM contributed to Fig. 3.1 and 3.2

Chandrasekhar A contributed to Fig. 3.2

58 3.1 Rationale

In a previous study, King et al. identified 11 IL1B-induced inflammatory mRNAs whose expression was not reduced by the translational blocker, cycloheximide, yet their repression by dexamethasone was cycloheximide-sensitive (273). These data may indicate that glucocorticoid- dependent gene expression is important in the repression of these inflammatory genes. In addition, a NR3C1 receptor antagonist and siRNA-mediated knock-down of NR3C1 attenuated the dexamethasone-dependent repression of these 11 IL1B-induced inflammatory mRNAs (273), further suggesting that the repression of these 11 IL1B-induced inflammatory mRNAs by glucocorticoid is NR3C1 dependent. In this regard, glucocorticoid-induced expression of DUSP1 plays an important role in the repression of MAPK signalling by glucocorticoids and may explain the gene expression-dependent repression by glucocorticoids (114, 291, 350). Indeed, DUSP1 expression is strongly, but transiently, induced by inflammatory stimuli and a clear role for DUSP1 in feedback inhibition of MAPKs and the expression of many inflammatory genes is established

(303, 304, 351-353). Similarly, DUSP1 is implicated in the glucocorticoid-dependent repression of CXCL1 and IL6 from human airway smooth muscle cells, and IL8 from human airway epithelial cells (354-356). However, glucocorticoid-dependent repression of inflammatory gene expression in DUSP1-/- mice was gene specific (281). For example, glucocorticoid-induced repression of

CSF2 and IL1A was strongly attenuated, but the repression of TNF, IL1B and PTGS2 was only partially attenuated (281). In addition, even though the expression of inflammatory gene expression often depends on MAPKs, a role for glucocorticoid-induced DUSP1 in the repression of inflammatory gene expression can be difficult to demonstrate (114, 283). Therefore, in the current chapter, possible repressive roles of DUSP1 that is induced by IL1B and IL1B plus the synthetic glucocorticoid, dexamethasone, is examined. Since pulmonary A549 cells, like primary

59 bronchial epithelial cells, show glucocorticoid-dependent repression of inflammatory gene expression, they have been selected for this study (273).

3.2 Hypothesis

The hypothesis being tested in this chapter is that glucocorticoid-induced DUSP1 may play an important role in the repression of inflammatory gene expression by glucocorticoids.

3.3 Results 3.3.1 Effect of dexamethasone on IL1B-induced MAPK activation and DUSP1 expression

Effect of IL1B, in the presence and absence of the synthetic glucocorticoid, dexamethasone, on the activation of p38, ERK and JNK MAPK pathways was examined at different time points. The phosphorylation, as assessed by T-x-Y motif, of the p38, ERK and JNK MAPK cascades was robustly activated by IL1B. Co-treatment with dexamethasone repressed all three MAPK pathways at 1, 2 and 6 h post-treatment (Fig. 3.1A). Even though the repression by dexamethasone mainly occurred at, and after, 1 h, the peak of IL1B-induced MAPK activity, at 30 min, was weakly affected by dexamethasone, and in fact the repressive effect of dexamethasone was variable at this time, with no repression in some analyses and partial repression in others. Thereafter, the repression of IL1B-induced MAPK activation by dexamethasone enhanced gradually with time and the repression was maximum at 6 h post-treatment.

Dexamethasone-induced repression of MAPKs was correlated with the enhanced expression of

DUSP1 protein. DUSP1 was strongly induced by IL1B that reached a peak at 1 h, before declining at 2 h, and reaching basal levels by 6 h post-treatment (Fig. 3.1B). Following IL1B plus dexamethasone-treatment, DUSP1 protein was induced within 30 min. In the case of dexamethasone alone, DUSP1 was induced at 1 h and this was further enhanced by co-treatment

60 with IL1B. A similar, but lesser, enhancing effect of IL1B on dexamethasone-induced DUSP1 expression was observed at 2 h. At 6 h, DUSP1 expression was predominantly dexamethasone- dependent with no significant enhancement by IL1B (Fig. 3.1B).

A time (h) ¼ ½ 1 2 6 Dex + + + + + + + + + + IL1B - - + + - - + + - - + + - - + + - - + + P-ERK P-p38 P-JNK

GAPDH B time (h) ¼ ½ 1 2 6 Dex + + + + + + + + + + IL1B - - + + - - + + - - + + - - + + - - + + DUSP1 GAPDH 2.0 *** 1.5 *** * 1.0 *** ** ** ** *** *** DUSP1 ***

/GAPDH 0.5 * *** * 0.0

Figure 3.1. Effect of dexamethasone and IL1B on MAPK phosphorylation and DUSP1 expression. A, A549 cells were either not stimulated or stimulated with IL1B (1 ng/ml), dexamethasone (Dex 1 μM) or a combination of the two as indicated. Cells were harvested after 0.15, 0.5, 1, 2 or 6 h and total proteins were prepared for western blot analysis of phospho-ERK (P-ERK), phospho-p38 (P-p38), phospho-JNK (P-JNK) and GAPDH. Blots representative of at least 4 such experiments are shown. B, As in A, cells were harvested at the times indicated for western blot analysis of DUSP1 and GAPDH. Following densitometric analysis, data (N = 9 - 11) were normalized to GAPDH and plotted as means ± S.E. Significance, using ANOVA with a Bonferroni's multiple comparison test is indicated. *, p < 0.05; **, p < 0.01; ***, p < 0.001.

61 3.3.2 Effect of dexamethasone on IL1B-induced gene expression

Treatment with IL1B enhanced the expression of 11 inflammatory mRNAs (official human genome nomenclature committee symbol with other commonly used gene symbols in brackets);

CCL2 (MCP1), CCL20 (MIP3α), CSF2 (GM-CSF), CXCL1 (GROα), CXCL2 (GROβ, MIP2α),

CXCL3 (GRO, MIP2β), IL6 (IL-6), IL8 (IL-8, CXCL8), ISG20 (CD25), OLR1 (LOX1) and

PTGS2 (COX-2), and these inflammatory genes were repressed by dexamethasone in a cycloheximide-sensitive manner (Fig. 3.2A & B) (273). IL1B-induced expression of CXCL1, IL8 and IL6 increased gradually over 6 h. Conversely, CXCL2 mRNA expression reached a peak at 1 h and thereafter declined by 6 h. In respect of CCL2, CCL20, CXCL3 and PTGS2 mRNAs, the expression was enhanced from 1 h to 2 h and maintained at this level for up to 6 h. conversely,

CSF2 mRNA was highly induced at 1 h, reached a maximum level at 2 h, and reduced by 6 h (Fig.

3.2A). Since these kinetics are consistent with a rapid induction of mRNA expression, these genes are considered immediate/early response genes whose expression was induced by IL1B and not affected by the translational blocker, cycloheximide (273).

In marked contrast, ISG20 and OLR1 mRNAs were not induced by IL1B at 1 h. At 2 h, there was only a modest increase, and at 6 h, the mRNA expression was substantially enhanced by IL1B

(Fig. 3.2B). Even though these kinetics were more consistent with late phase/secondary response genes, the expression of these mRNAs was insensitive to cycloheximide in the presence of IL1B

(114). Nevertheless, because the timing of induction for these two mRNAs was different from the nine genes showing classical primary response kinetics, they were simply referred to as delayed or secondary response genes.

62 A 2 2 4 CCL2 CCL20 CSF2 1 1 2 *** ** ** *** *** 0 * 0 0 * *** 2 1.5 1.5 CXCL1 CXCL2 CXCL3 NS 1.0 1.0 1 IL1B 0.5 0.5 *** *** *** ** IL1B + Dex ** ** *** 0 0.0 0.0

gene/GAPDH 1.5 2 4 IL6 IL8 PTGS2 1.0 2 1 0.5 *** *** **** *** *** 0 *** 0.0 ** 0 *** 0 2 4 6 0 2 4 6 0 2 4 6 time (h) time (h) time (h) B C 6 ISG20 100 N = 9 genes CCL2 4 100 CCL20 *** 2 CSF2 *** ** *** CXCL1

+Dex 0 +Dex

at each time) at eachtime)

 CXCL2 50  ***

  50 3 OLR1 CXCL3

IL-1 IL6 IL-1 2 gene/GAPDH IL8 *** 1 PTGS2 IL-1 of %

(% of(% IL-1 *** ( ** 0 0 0 0 2 4 6 0 2 4 6 1 2 6 time (h) time (h) time (h) D 200 30000 CSF2 CXCL1 150 20000 100 time (h) 1 2 6 10000 *** Dex + + + + + + 50 *** *** *** IL1B - - + + - - + + - - + + 0 0 1000 4000 PTGS2

(pg/ml) Release IL6 IL8 750 3000 GAPDH 500 2000 250 1000 * *** *** 0 *** 0 *** *** 0 2 4 6 0 2 4 6 time (h) time (h)

Figure 3.2. Effect of dexamethasone and IL1B on inflammatory gene expression. A & B, A549 cells were either not stimulated (NS) () or stimulated with IL1B (1 ng/ml) () or IL1B plus dexamethasone (1 μM) (IL1B + Dex) () as indicated. Cells were harvested after 1, 2 or 6 h for real-time PCR analysis of the indicated genes and GAPDH. Data (N = 11) were normalized to GAPDH and are plotted as means ± S.E. Significance was tested relative to time-matched IL1B- treated samples using ANOVA with a Bonferroni post test. *, p < 0.05; **, p < 0.01; ***, p < 0.001. Genes were grouped according to expression pattern with apparent ‘early-phase’ mRNAs in A and ‘delayed response’ mRNAs in B. C, The effect of IL1B + dexamethasone for each inflammatory mRNA in panel A is plotted as a percentage of IL1B for each time (left panel). The overall effect of dexamethasone in the presence of IL1B for the 9 genes combined is plotted as means ± S.E. (right panel). D, Cells were treated as in A and the supernatants harvested after 1, 2 or 6 h for cytokine/chemokine release measurement. Data (N = 6) expressed as pg/ml are plotted as means ± S.E. (left panel). Significance, relative to IL1B-treated samples was tested using ANOVA with a Bonferroni post test. *, p < 0.05; **, p < 0.01; ***, p < 0.001. Cells were harvested at the times indicated for western blot analysis of PTGS2 and GAPDH (right panel). Blots representative of at least 4 such experiments are shown.

63 The mRNA expression of all IL1B-induced 11 inflammatory genes was significantly repressed by dexamethasone at 6 h (Fig. 3.2A & B). This effect was more modest at 1 and 2 h, and CCL20 showed no repression at 1 h. Dexamethasone-dependent repression of the 9 primary response genes (Fig. 3.2A) increased progressively with time (Fig. 3.2C). However, this was not the case with respect to ISG20 and OLR1, where there was a little mRNA accumulation at 1 or 2 h. These findings were consistent with the previous data showing that the repression of these 11 mRNAs was prevented or reduced by cycloheximide, and the fact that time is necessary for the synthesis of the glucocorticoid-induced gene-products mediating repression (114).

The expression of CSF2, CXCL1, IL6 and IL8 released into the supernatant along with the cellular expression of PTGS2 was also analysed. In each case, there was a time-dependent enhancement by IL1B and dexamethasone co-treatment produced a significant repression at all times (Fig.

3.2D).

3.3.3 Effect of MAPK inhibitors on IL1B-induced gene expression

To explore the role of p38, ERK and JNK MAPK pathways in the expression of IL1B-induced 11 inflammatory mRNAs, the p38 inhibitor, SB203580, and the MAPK/ERK kinase (MEK) 1/2 inhibitor, U0126, were tested along with JNK inhibitor, JNK inhibitor 8. JNK inhibitor 8 is a recently identified, highly selective, cell permeable inhibitor, of the JNK MAPK pathway with a potency of ~1 µM on cellular JUN phosphorylation (357). A549 cells were treated with increasing concentration of JNK inhibitor 8 prior to stimulation with IL1B and western blot analysis for phosphorylation of cellular JUN was carried out. Treatment with IL1B significantly increased the phosphorylation of JUN and this was prevented by JNK inhibitor 8 (EC50 = 0.8 µM). Near maximal

64 (83%) repression was achieved at 10 µM, and hence, this concentration was used for subsequent analyses (Fig. 3.3).

time (h) ½ JNK-IN-8 IL1B - + + + + + + + P-JUN JUN

100

IL1B) 50

JUN/JUN (%

- ***

P *** 0 *** -7 -6 -5 -4 NSIL-1 Log [JNK-IN-8 (M)]

Figure 3.3 Effect of JNK inhibitor 8 on IL1B-induced JUN phosphorylation. A549 cells were either not stimulated (NS), treated with IL1B (1 ng/ml) or pre-treated with increasing concentrations of JNK inhibitor 8 (JNK-IN-8) for 30 min prior to IL1B stimulation. Cells were harvested after 0.5 h and total proteins prepared for western blot analysis of phospho-JUN (P-JUN) and JUN. Representative blots are shown. Following densitometric analysis, data (N = 5), normalized to JUN, were expressed as a percentage of IL1B and are plotted as mean ± S.E. Significance, using ANOVA with a Dunnett's post test is indicated. ***, p < 0.001.

Similar to JNK inhibitor 8, SB203580 and UO126 were also used at a maximally effective concentration (10 µM) as determined previously by analysis of substrate phosphorylation and functional responses in A549 cells (358, 359). SB203580 produced a partial and time-dependent inhibition of most of the IL1B-induced mRNAs (Fig. 3.4). However, while IL1B-induced CXCL2 and ISG20 were unaffected, IL6 was strongly repressed following SB203580 treatment. In contrast, UO126 and JNK inhibitor 8 produced moderate to weak repressive effects (Fig. 3.4).

65 time (h) 1 2 6 1 2 6 1 2 6 JNK-IN-8 + + + + + + + + + UO126 + + + + + + + + + SB203580 + + + + + + + + + IL1B -++++ -++++ -++++ -++++ -++++ -++++ -++++ -++++ -++++ 8 8 CCL2 CCL20 2.0 CSF2 6 6 1.5 * * 4 * 4 * 1.0 * 2 2 0.5 * * * * 0 0 * 0.0

CXCL1 4 CXCL2 3 CXCL3 4 3 2 * * * 2 * 2 * * * 1 1 * gene/GAPDH 0 0 0 IL6 1.5 IL8 2.0 PTGS2 2 * * 1.5 1.0 * * ** * * ** * ** ** * * 1 * * ** * 1.0 ** * * ** 0.5 * * * * * * ** ** 0.5 * *** * * * 0 * 0.0 0.0

time (h) 1 2 6 1 2 6 JNK-IN-8 + + + + + + UO126 + + + + + + SB203580 + + + + + + IL1B -++++ -++++ -++++ -++++ -++++ -++++ 8 8 ISG20 OLR1 6 6 4 4 ** gene/ * ** GAPDH 2 * 2 * * ** * ** 0 0 *** Figure 3.4 Effect of MAPK inhibitors on the mRNA expression of inflammatory genes. A549 cells were either not stimulated, treated with IL1B (1 ng/ml) or pre-treated with UO126, SB203580 or JNK inhibitor 8 (JNK-IN-8) each at 10 µM for 30 min prior to IL1B stimulation. Cells were harvested after 1, 2 or 6 h for real-time PCR analysis of the indicated genes and GAPDH. Data (N = 8) were normalized to GAPDH and plotted as means ± S.E. ‘Early-phase’ genes are shown in top panel and ‘delayed response’ genes in bottom panel. Significance, relative to time-matched IL1B-treated samples, was tested by ANOVA with a Dunnett's post test. *, p < 0.05; **, p < 0.01; ***, p < 0.001.

66 At the level of cytokine production, IL1B-induced release of CSF2, CXCL1, IL6 and IL8 was significantly reduced by SB203580 at 2 and 6 h post-treatment (Fig. 3.5). Even though the release of CSF2 and CXCL1 at 2 and 6 h was robustly inhibited, the release of IL6 and IL8 was only partially affected by UO126 (Fig. 3.5). JNK inhibitor 8 also produced a significant attenuation of

CSF2 and CXCL1 release, but showed a partial, non-significant, inhibition of IL6 and IL8 at either time point (Fig. 3.5).

time (h) 2 6 JNK-IN-8 + + UO126 + + SB203580 + + IL1B -++++ -++++ 90 10 60 5

***

**

*** CSF2 30

***

***

*** 0 0 5000 100 2500

*** 50 ***

***

***

***

***

CXCL1 0 0

(pg/ml) 15 450 Release Release

10 300 *

IL6 5 ** 150

*** 0 *** 0 1500 10000

**

1000 *

**

IL8 5000 500 * 0 0 Figure 3.5 Effect of MAPK inhibitors on the protein expression of inflammatory genes. A549 cells were either not stimulated, treated with IL1B (1 ng/ml) or pre-treated with UO126, SB203580 or JNK inhibitor 8 (JNK-IN-8) each at 10 µM for 30 min prior to IL1B stimulation. Supernatants were harvested after 2 or 6 h for measurement of cytokine/chemokine release. Data (N = 8) expressed in pg/ml are plotted as means ± S.E. Significance, relative to time-matched IL1B-treated samples, was tested by ANOVA with a Dunnett's post test. *, p < 0.05; **, p < 0.01; ***, p < 0.001

Inflammatory stimuli-induced DUSP1 expression is, in part, dependent on MAPK pathways, and inhibition of a single MAPK pathway may reduce or prevent DUSP1 expression (112, 286, 360).

Since DUSP1 inhibits all three MAPK pathways, inhibition of any one MAPK pathway could

67 correspondingly enhance the activation of the remaining two MAPK pathways. Equally, cross- feedback control by p38 MAPK down-regulates the ERK pathway, which may therefore be enhanced in the presence of a p38 inhibitor (361). Furthermore, dexamethasone also produces the repression of all three MAPKs together (Fig. 3.1A). Thus, the effect of simultaneous inhibition of three MAPK pathways was investigated on inflammatory gene expression (Fig. 3.6). The expression of the 9 primary response genes (shown in Fig. 3.2A) was strongly induced by IL1B and, in each case, this was significantly and substantially inhibited by the combined MAPK inhibitors (Fig. 3.6A). A similar, almost complete, repression was also observed for ISG20 and

OLR1 (Fig. 3.6B). Equally, IL1B-induced release of CSF2, CXCL1, IL6 and IL8 was also completely repressed by MAPK inhibitors in combination (Fig. 3.6C & D). In the case of PTGS2 protein expression, it was weakly induced by IL1B at 2 h and moderately reduced by the individual kinase inhibitors (Fig. 3.6D). By 6 h, there was robust PTGS2 expression and this was significantly attenuated by each MAPK inhibitor. In contrast, IL1B-induced PTGS2 expression, at 2 and 6 h, was almost completely blocked by the combined MAPK inhibitors (Fig. 3.6D). The MAPK inhibitors, alone or in combination, had no effect on cell viability as measured by MTT assay

(Table 3.1).

68 A 15 1.5 3 CCL2 CCL20 CSF2 10 1.0 2 5 *** 0.5 1 *** *** ** 0 0.0 0 ** * 6 CXCL1 1.2 CXCL2 1.2 CXCL3 NS 4 0.8 0.8 IL1B 2 0.4 0.4 *** IL1B + UO *** *** *** *** ** *** 0 0.0 0.0 + SB + J8

gene/GAPDH 1.5 1.5 2.0 IL6 IL8 PTGS2 1.5 1.0 1.0 1.0 0.5 0.5 0.5 * ** ** ** *** 0.0 ** *** 0.0 0.0 0 2 4 6 0 2 4 6 0 2 4 6 time (h) time (h) time (h)

B 10 15 8 ISG20 OLR1 6 10 4 5 gene/ 2 *** GAPDH *** 0 0 0 2 4 6 0 2 4 6 time (h) time (h) C D time (h) 2 6 150 5000 CSF2 CXCL1 JNK-IN-8 + + + + 100 UO126 + + + + 2500 50 SB203580 + + + + *** *** - + + + + + - + + + + + 0 ** 0 IL1B 200 15000 PTGS2

(pg/ml) Release 150 IL6 IL8 10000 GAPDH 100 50 5000 0.3 *** 0 0 *** 0.2 * 0.1 ** 0 2 4 6 0 2 4 6 *** ** ** time (h) time (h) PTGS2 0.0 *** /GAPDH Figure 3.6 Effect of MAPK inhibitors on IL1B-induced inflammatory gene expression. A & B, A549 cells were either not stimulated (NS) (), treated with IL1B (1 ng/ml) () or pre-treated with a combination of UO126, SB203580 plus JNK inhibitor 8 each at 10 µM (UO+SB+J8) for 30 min prior to IL1B stimulation (). Cells were harvested after 1, 2 or 6 h for real-time PCR analysis of the indicated genes and GAPDH. Data (N = 4) were normalized to GAPDH and plotted as means ± S.E. ‘Early-phase’ genes are shown in A and ‘delayed response’ genes in B. C, Cells were treated as in A and the supernatants harvested after 1, 2 or 6 h for the measurement of cytokine/chemokine release. Data (N = 4) expressed in pg/ml are plotted as means ± S.E. D, Cells were treated as in A and harvested at the times indicated for western blot analysis of PTGS2 and GAPDH. Representative blots are shown. Following densitometric analysis, data (N = 4), normalized to GAPDH, are plotted as means ± S.E. Significance, relative to time-matched IL1B- treated samples, was tested by ANOVA with a Dunnett's post test. *, p < 0.05; **, p < 0.01; ***, p < 0.001.

69 Table 3.1 Effect of MAPK inhibitors on cell viability.

A549 cells were either not treated, treated with IL1B (1 ng/ml), or pre-treated with 10 μM of either SB203580, UO126, JNK inhibitor 8 or a combination of SB203580, UO126 plus JNK inhibitor 8 each at 10 µM (UO+SB+J8) for 30 min prior to IL1B stimulation for 1, 2 or 6 h. Cell viability was determined by MTT assay. Data (N = 4), expressed as OD584, are showed as means ± S.E. Significance, versus IL1B, using ANOVA with a Dunnett's post test is indicated. However, significance was not achieved for any of the tested samples.

Time (h) Treatment OD value ± S.E. No treatment 0.597 ± 0.1260 IL1B 0.684 ± 0.1738 IL1B + SB203580 0.589 ± 0.1275 1 h IL1B + UO126 0.622 ± 0.1387 IL1B + JNK8 0.614 ± 0.0939 IL1B + UO + SB + J8 0.556 ± 0.1287 No treatment 0.533 ± 0.1111 IL1B 0.531 ± 0.0892 IL1B + SB203580 0.595 ± 0.1263 2 h IL1B + UO126 0.628 ± 0.1483 IL1B + JNK8 0.637 ± 0.1733 IL1B + UO + SB + J8 0.498 ± 0.1108 No treatment 0.485 ± 0.1238 IL1B 0.604 ± 0.1606 IL1B + SB203580 0.525 ± 0.0980 6 h IL1B + UO126 0.484 ± 0.1295 IL1B + JNK8 0.528 ± 0.1084 IL1B + UO + SB + J8 0.523 ± 0.0918

3.3.4 Effect of DUSP1 over-expression on MAPK activation and inflammatory gene expression

DUSP1 over-expression was confirmed by western blotting. Control GFP adenovirus had no effect on DUSP1 expression. IL1B-induced phosphorylation of ERK and p38 was substantially reduced following adenoviral over-expression of DUSP1 (Fig. 3.7A), and was consistent with previous data (114, 283). IL1B-induced JNK phosphorylation was also inhibited following DUSP1 over- expression and, in each case, GFP expressing adenovirus had no effect (Fig. 3.7A).

70 A time (h) 1 Ad-DUSP1 + + Ad-GFP + + IL1B - - - + + + DUSP1 P-ERK P-p38

P-JNK

GAPDH

B time (h) 1 2 6 C time (h) 1 2 6 Ad-DUSP1 + + + + + + Ad-DUSP1 + + + + + + Ad-GFP + + + + + + Ad-GFP + + + + + + IL1B - - - + + + - - - + + + - - - + + + IL1B - - - + + + - - - + + + - - - + + + 1.0 1.0 1.0 * 8 8 8 * * 6 6 6 0.5 0.5 0.5 4 4 4

CCL2 ISG20 2 2 2 0.0 0.0 0.0 0 0 0 * * 1.0 1.0 * 1.0 8 8 8 * 6 6 6

0.5 0.5 0.5 Gene/GAPDH 4 4 4

OLR1 CCL20 2 2 2 0.0 0.0 0.0 0 0 0 1.5 1.5 * 1.5 * 1.0 1.0 1.0 D time (h) 2 6

CSF2 0.5 0.5 0.5 * * Ad-DUSP1 + + + + 0.0 0.0 0.0 Ad-GFP + + + + 2 2 2 ** * ** IL1B - - - + + + - - - + + + * *** 1 1 1 60 100 *** *** CXCL1 ** 40 * 0 0 0 * 50 CSF2 20 ** ** 0 0 1.0 ** 1.0 1.0 * *** ** 400 *** 0.5 0.5 0.5 ** 4000 CXCL2 300 **

0.0 0.0 0.0 200 ** 2000 Gene/GAPDH * CXCL1 100 1.0 * 1.0 0 0 1.0 ** ** ** 60 1000 *** 0.5 * 0.5 0.5 ***

CXCL3 40 Release (pg/ml)Release 0.0 0.0 0.0 IL6 500 *** 20 ** 2 2 *** *** *** 2 0 0

IL6 1 ** 1 1 1500 6000 ** * ** 1000 4000 0 0 0 ***

IL8 *** 1.5 1.5 * 1.5 500 2000 *** 1.0 ** 1.0 1.0 *** 0 0 IL8 ** 0.5 0.5 0.5 time (h) 2 6 0.0 0.0 0.0 Ad-DUSP1 + + + + ** Ad-GFP + + + + ** 2 2 2 * - - - + + + * IL1B - - - + + + 1 * 1 1 PTGS2

PTGS2 0 0 0 GAPDH

Figure 3.7 Effect of DUSP1 over-expression on IL1B-induced MAPK phosphorylation and inflammatory gene expression.

71 Figure 3.7 continued. A, A549 cells were either not infected or infected with Ad5-DUSP1, or Ad5-GFP at a MOI of 10 for 24 h before IL1B treatment (1 ng/ml). After 1 h, the cells were harvested for western blot analysis of DUSP1, phospho-ERK (P-ERK), phospho-p38 (P-p38), phospho-JNK (P-JNK) and GAPDH. Blots representative of at least 4 such experiments are shown. B & C, Cells were treated as in A and harvested at 1, 2 or 6 h for real-time PCR analysis of the indicated genes and GAPDH. Data (N = 4) were normalized to GAPDH and plotted as means ± S.E. Significance relative to time-matched IL1B and Ad5-GFP-treated samples, was tested by ANOVA with a Bonferroni post test. *, p < 0.05; **, p < 0.01; ***, p < 0.001. ‘Early-phase’ genes are shown in B and ‘delayed response’ genes in C. D, Cells were treated as in A and the supernatants harvested after 2 or 6 h for cytokine/chemokine release measurement. Data (N = 4) expressed in pg/ml are plotted as means ± S.E. (top panel). Significance, relative to time-matched IL1B and Ad5-GFP-treated samples, was tested by ANOVA with a Bonferroni post test. *, p < 0.05; **, p < 0.01; ***, p < 0.001. Cells were also harvested at the times indicated for western blot analysis of PTGS2 and GAPDH (lower panel). Blots representative of at least 4 such experiments are shown.

In respect of the 11 IL1B-induced genes, DUSP1 over-expression produced a significant inhibition

(Fig. 3.7B&C). Equally, the protein expression for CSF2, CXCL1, IL6, IL8 and PTGS2 was also significantly attenuated following DUSP1 over-expression (Fig. 3.7D). The control GFP adenovirus had no effect. Thus, like dexamethasone, DUSP1 over-expression also produced a significant inhibition of all three MAPK pathways and IL1B-induced inflammatory gene expression.

3.3.5 Effect of DUSP1 siRNA on MAPK activation by IL1B in the absence and presence of dexamethasone

The control siRNA targeted to LMNA had no effect on DUSP1 expression at any time (Fig. 3.8A).

Conversely, IL1B or IL1B plus dexamethasone-induced DUSP1 was substantially and significantly inhibited by two separate DUSP1 targeting siRNAs at all times tested (Fig. 3.8B). As described above, IL1B activated all three MAPK pathways and dexamethasone produced a significant repression of MAPKs at 1, 2 and 6 (Fig. 3.9). LMNA siRNA had no effect on MAPK activation induced by IL1B or IL1B plus dexamethasone (Fig. 3.9A, Table 3.2).

72 A time (h) 1 2 6 LMNA siRNA + + + + + + Dex + + + + + + IL1B - + + + + - + + + + - + + + + DUSP1 GAPDH 2 * ** *

1 ***

DUSP1 /GAPDH 0 B time (h) 1 2 6 DUSP1 siRNA 2 + + + + + + DUSP1 siRNA 1 + + + + + + LMNA siRNA + + + + + + Dex + + + + + + + + + IL1B - + + + + + + - + + + + + + - + + + + + + DUSP1 GAPDH 600 *** *** 400 *** *** * *** *** ** * **

200 * * * * (% IL1B IL1B (%+

lamin siRNA) lamin 0

DUSP1/GAPDH Figure 3.8 Effect of LMNA- and DUSP1-targeting siRNAs on IL1B and IL1B plus dexamethasone-induced DUSP1 expression. A549 cells were: A, incubated without or with LMNA-specific siRNA; or B, incubated with LMNA (control) or DUSP1-specific siRNAs. After 24 h, cells were treated with IL1B (1 ng/ml) or IL1B plus dexamethasone (1 μM) (Dex) as indicated. Cells were harvested at 1, 2 or 6 h for western blot analysis of DUSP1 and GAPDH. Representative blots are shown. Following densitometric analysis, data in A (N = 7) were normalized to GAPDH and are plotted as mean ± S.E. In B, data (N = 10), normalized to GAPDH, were expressed as a percentage of LMNA siRNA plus IL1B-stimulated cells for each time and plotted as means ± S.E. In each case, significance was tested using ANOVA with a Bonferroni's multiple comparison test is indicated. *, p < 0.05; **, p < 0.01; ***, p < 0.001.

However, in the presence of DUSP1 targeting siRNAs, 1 h post-IL1B treatment, the phosphorylation of all three MAPKs was significantly enhanced relative to LMNA control (Fig.

3.9 B & C). This effect was not seen at 2 or 6 h post-IL1B treatment. Thus, DUSP1 plays a significant role in the feedback control of MAPKs at 1 h, but not at later times.

73 IL1B-induced phosphorylation of ERK, p38 and JNK was significantly reduced by dexamethasone co-treatment at each time (Fig. 3.9A-C). In the presence of two DUSP1 targeting siRNAs the phosphorylation of MAPKs was significantly enhanced at 1 h relative to the IL1B plus dexamethasone-treated control (Fig. 3.9B & C). Thus, IL1B plus dexamethasone-induced DUSP1 reduced MAPK activation. However, while evaluating the role of DUSP1 in dexamethasone- induced repression, it is essential to consider the inhibitory feedback role of DUSP1. Thus, in Fig.

3.9D, the effect of IL1B plus dexamethasone on MAPK phosphorylation was expressed as a percentage of the IL1B-induced MAPK phosphorylation for each siRNA treatment. In this way the repressive effect of dexamethasone can be assessed and compared in the presence and absence of DUSP1.

In the presence of dexamethasone, the phosphorylation of each MAPK was reduced to ~50% of the IL1B plus LMNA control siRNA-treated level at 1 h (Fig. 3.9C & D). However, this repressive effect was significantly and substantially attenuated by the DUSP1-targeting siRNAs (Fig. 3.9D)

(i.e. the percentage MAPK phosphorylation for IL1B plus dexamethasone/IL1B was close to

100%), and further suggested that dexamethasone-dependent repression of IL1B-induced MAPK phosphorylation was prevented by DUSP1 knock-down. Therefore, these data confirm that DUSP1 not only plays a key role in the feedback control of ERK, p38 and JNK MAPKs, but is also responsible for their repression by dexamethasone at 1 h.

74 A B time (h) 1 2 6 DUSP1 siRNA 2 + + + + + + time (h) 1 2 6 DUSP1 siRNA 1 + + + + + + LMNA siRNA + + + + + + LMNA siRNA + + + + + + Dex + + + + + + Dex + + + + + + + + + IL1B - + + + + - + + + + - + + + + IL1B - + + + + + + - + + + + + + - + + + + + + P-ERK P-ERK P-p38 P-p38

P-JNK P-JNK

GAPDH GAPDH

C time (h) 1 2 6 D DUSP1 siRNA 2 + + + + + + time (h) 1 2 6 DUSP1 siRNA 1 + + + + + + DUSP1 siRNA 2 + + + LMNA siRNA + + + + + + DUSP1 siRNA 1 + + + Dex + + + + + + + + + LMNA siRNA + + + IL1B - + + + + + + - + + + + + + - + + + + + + IL1B+Dex + + + + + + + + + 300 * ** ** 200 *** 100 * 100 100 *** * * 100 50 50 50

P-ERK

P-ERK * * * A) 0 0 0 0 150 200 * * * * ** 100 100 *** ** 100 * 100 50 50 P-p38 P-p38 50 * * * 0 0 0 0 P-MAPK/GAPDH 300 * 200 * 150 * **

(% IL1B + LMNA siRNA) + LMNA (% IL1B

* IL1Brelevant + % siRN * Effect ofon Dex P-MAPK 150 100 200 ** *** ( *** 100 100

P-JNK 100 50 * * P-JNK 50 50 ** 0 0 0 0

Figure 3.9 Effect of LMNA- and DUSP1-targeting siRNA on IL1B-induced MAPK phosphorylation. A549 cells were; A, incubated without or with LMNA-specific siRNA, or B, incubated with LMNA (control) or DUSP1-specific siRNAs. After 24 h, cells were treated with IL1B (1 ng/ml) or IL1B plus dexamethasone (1 μM) (Dex) as indicated. Cells were then harvested at 1, 2 or 6 h and total proteins were prepared for western blot analysis of phospho-ERK (P-ERK), phospho-p38 (P-p38), phospho-JNK (P-JNK) and GAPDH. Representative blots are shown. Blots representative of at least 6 - 9 such experiments are shown. Densitometric data from A appear in Table 3.2 shown below. C, Following densitometric analysis, data from B were normalized to GAPDH, expressed as a percentage of LMNA siRNA plus IL1B-stimulated for each time and plotted as means ± S.E. Significance was tested using ANOVA with a Newman-Keul multiple comparison test. Significance between: LMNA control siRNA plus IL1B and each of the DUSP1 targeting siRNAs plus IL1B, and the LMNA control plus IL1B plus Dex is shown. Other comparisons are specifically indicated. *, p < 0.05; **, p < 0.01; ***, p < 0.001. D, For each P- MAPK at each time, the effect of IL1B plus dexamethasone expressed as a percentage of IL1B for each of the three individual siRNAs is plotted as a mean ± S.E. The percent IL1B plus dexamethasone/IL1B for the LMNA siRNA is compared with that for each DUSP1-specific siRNAs using ANOVA with a Dunnett's post test. *, p < 0.05; **, p < 0.01; ***, p < 0.001.

75 These data contrast significantly with effects of IL1B treatment observed at 6 h (Fig. 3.9B-D).

Even though IL1B-induced MAPK activation was reduced relative to earlier times, dexamethasone still produced a significant repression at 6 h (Fig. 3.9A & B). However, this was not affected by the DUSP1-targeting siRNAs, suggesting that the repressive effects of dexamethasone observed at 6 h were not due to DUSP1 (Fig. 3.9B-D). Conversely, the effect of the two DUSP1 siRNAs at

2 h post-IL1B plus dexamethasone treatment was intermediate between the effects at 1 and 6 h.

The phosphorylation of MAPKs induced by IL1B at 2 h was increased, but did not reach significance (Fig. 3.9C). Similarly, there was non-significant losses of dexamethasone-dependent repression following DUSP1 knock-down at 2 h (Fig. 3.9D).

Table 3.2 Densitometry analysis for the effect of LMNA targeting siRNA (LsiRNA) on MAPK phosphorylation.

MAPKs/GAPDH Time Gene Symbol NS IL1B IL1B + LSiRNA IL1B + Dex IL1B + Dex + LsiRNA P-ERK 0.47 ± 0.15 0.83 ± 0.21* 0.81 ± 0.20 0.44 ± 0.13** 0.49 ± 0.12* 1h P-p38 0.086 ± 0.028 0.61 ± 0.11*** 0.69 ± 0.14 0.33 ± 0.086** 0.42 ± 0.10* P-JNK 0.16 ± 0.067 0.52 ± 0.10*** 0.52 ± 0.069 0.25 ± 0.053*** 0.32 ± 0.046** P-ERK 0.39 ± 0.087 0.99 ± 0.24** 1.1 ± 0.29 0.44 ± 0.088* 0.55 ± 0.12* 2h P-p38 0.29 ± 0.14 0.90 ± 0.20** 0.85 ± 0.22 0.41 ± 0.093** 0.45 ± 0.098* P-JNK 0.16 ± 0.029 0.57 ± 0.20** 0.50 ± 0.12 0.24 ± 0.045*** 0.17 ± 0.030** P-ERK 0.60 ± 0.19 1.3 ± 0.49* 1.4 ± 0.45 0.59 ± 0.15* 0.72 ± 0.18* 6h P-p38 0.19 ± 0.11 0.71 ± 0.24*** 0.73 ± 0.25 0.21 ± 0.060** 0.29 ± 0.15 P-JNK 0.24 ± 0.11 0.44 ± 0.056** 0.43 ± 0.083 0.18 ± 0.052*** 0.24 ± 0.061**

3.3.6 Effect of DUSP1 knock-down on IL1B-induced inflammatory gene mRNA expression in the absence and presence of dexamethasone.

The expression of all 11 inflammatory mRNAs was induced by IL1B and this was significantly repressed by dexamethasone. The LMNA siRNA had no effect on IL1B and IL1B plus dexamethasone-induced gene expression at any time (Table 3.3). Therefore, in subsequent analyses, the effects of DUSP1-targeting siRNAs were compared solely to the LMNA control siRNA. The expression of all 9 acute phase mRNAs (shown in Fig. 3.2A) was increased by DUSP1

76 knock-down at 1 h following IL1B treatment (Fig. 3.10A). This effect was significant in respect of one or both DUSP1-targeting siRNAs for CCL2, CSF2, CXCL3, IL6, IL8 and PTGS2.

However, by 2 h, with the exception of IL6, this trend was markedly reduced or not seen post-

IL1B treatment. Conversely, the expression of all 9 acute phase mRNAs produced a general trend toward attenuated expression by DUSP1 knock-down at 6 h post-IL1B treatment. This was significant for CCL2, CCL20, CSF2, CXCL1, CXCL2, CXCL3 and PTGS2 in respect of one or both DUSP1 siRNAs (Fig. 3.10A). IL1B-induced expression of the two delayed response genes,

ISG20 and OLR1, was also enhanced by DUSP1 knock-down (Fig. 3.10C). This was significant for OLR1 at 1 h for both siRNAs and for one siRNA at 2 h. There was no effect of the DUSP1 siRNAs on either gene at 6 h (Fig. 3.10C).

Dexamethasone co-treatment produced a significant repression of IL1B-induced expression of all mRNAs at 2 and 6 h (Fig. 3.10A). IL1B-induced all the mRNAs, with the exception of CCL20, were also repressed by dexamethasone at 1 h post-treatment (Fig. 3.10A). Although this was not significant for CSF2, these data were consistent with Fig. 3.2. In respect of DUSP1-targeting siRNAs, there was a general trend towards higher levels of inflammatory gene expression at 1 h post-treatment. Indeed, the mRNA expression of CCL2, CCL20, CSF2, CXCL1, CXCL2, CXCL3,

IL8 and PTGS2 was enhanced, with at least one DUSP1 siRNA, relative to IL1B plus dexamethasone (plus LMNA siRNA) treated cells (Fig. 3.10A). Even though, ISG20 was not induced by IL1B at 1 h, dexamethasone produced a significant repression of ISG20 and this was inhibited by the DUSP1 siRNAs (Fig. 3.10C). IL1B-induced OLR1 expression was also repressed by dexamethasone, but this was unaffected by DUSP1 knock-down (Fig. 3.10C).

77 Table 3.3 Effect of LMNA targeting siRNA (LsiRNA) on inflammatory gene expression. A549 cells were incubated with LMNA-specific siRNA for 24 h before being treated with IL1B (1 ng/ml) or IL1B plus dexamethasone (1 μM) (Dex) as indicated. Cells were harvested after 1, 2 or 6 h for real-time PCR analysis of the indicated genes and GAPDH. Data (N = 11) were normalized to GAPDH and presented as means ± S.E. Significance, relative to time-matched IL1B and IL1B plus LMNA siRNA-treated samples, was tested by ANOVA with a Bonferroni's post test *, p < 0.05; **, p < 0.01; ***, p < 0.001.

Gene Symbol Gene/GAPDH Time NS IL1B IL1B + LSiRNA IL1B + Dex IL1B + Dex + LsiRNA CCL2 0.012 ± 0.0025 0.20 ± 0.042 0.14 ± 0.027 0.11 ± 0.018*** 0.086 ± 0.014*** CCL20 0.0013 ± 0.00041 0.31 ± 0.093 0.26 ± 0.072 0.30 ± 0.074 0.22 ± 0.045 CSF2 0.0014 ± 0.00040 0.46 ± 0.23 0.43 ± 0.21 0.18 ± 0.065 0.24 ± 0.10 CXCL1 0.0072 ± 0.0022 0.27 ± 0.049 0.65 ± 0.060 0.13 ± 0.017*** 0.47 ± 0.083*** CXCL2 0.019 ± 0.0050 0.97 ± 0.20 0.90 ± 0.18 0.85 ± 0.19 0.74 ± 0.18*** 1h CXCL3 0.012 ± 0.0031 0.51 ± 0.079 0.41 ± 0.068 0.32 ± 0.051*** 0.28 ± 0.041*** IL6 0.0033 ± 0.0017 0.78 ± 0.17 0.74 ± 0.16 0.34 ± 0.11*** 0.28 ± 0.090*** IL8 0.0083 ± 0.0029 0.50 ± 0.067 0.45 ± 0.062 0.29 ± 0.042*** 0.26 ± 0.035*** ISG20 0.46 ± 0.089 0.50 ± 0.094 0.48 ± 0.091 0.37 ± 0.070* 0.37 ± 0.082** OLR1 0.016 ± 0.0033 0.082 ± 0.016 0.058 ± 0.0098 0.039 ± 0.0075*** 0.034 ± 0.0060*** PTGS2 0.11 ± 0.023 0.55 ± 0.087 0.45 ± 0.074 0.31 ± 0.048*** 0.23 ± 0.044*** CCL2 0.013 ± 0.0025 1.0 ± 0.14 0.94 ± 0.11 0.46 ± 0.095*** 0.32 ± 0.025*** CCL20 0.044 ± 0.041 0.97 ± 0.12 0.98 ± 0.10 0.68 ± 0.11*** 0.68 ± 0.065*** CSF2 0.0039 ± 0.0023 3.0 ± 0.82 3.3 ± 0.94 1.1 ± 0.31*** 1.2 ± 0.35*** CXCL1 0.0061 ± 0.0016 0.72 ± 0.12 0.57 ± 0.087 0.30 ± 0.050*** 0.38 ± 0.085*** CXCL2 0.013 ± 0.0054 0.91 ± 0.19 0.85 ± 0.18 0.67 ± 0.18 0.62 ± 0.16 2h CXCL3 0.011 ± 0.0058 1.1 ± 0.13 0.95 ± 0.085 0.43 ± 0.068*** 0.35 ± 0.059*** IL6 0.0050 ± 0.0017 1.6 ± 0.45 1.4 ± 0.42 0.39 ± 0.15*** 0.27 ± 0.11*** IL8 0.0074 ± 0.0044 0.96 ± 0.12 0.81 ± 0.091 0.37 ± 0.055*** 0.37 ± 0.050*** ISG20 0.48 ± 0.096 0.78 ± 0.17 0.78+ ± 0.17 0.55 ± 0.15*** 0.50 ± 0.096*** OLR1 0.024 ± 0.0051 0.72 ± 0.12 0.55 ± 0.099 0.19 ± 0.037*** 0.15 ± 0.028*** PTGS2 0.081 ± 0.018 1.1 ± 0.16 0.99 ± 0.14 0.44 ± 0.081*** 0.29 ± 0.030*** CCL2 0.0084 ± 0.0012 1.3 ± 0.12 1.3 ± 0.20 0.099 ± 0.014*** 0.067 ± 0.011*** CCL20 0.00083 ± 0.000421.1 ± 0.23 1.1 ± 0.24 0.18 ± 0.037*** 0.19 ± 0.036*** CSF2 0.0017 ± 0.00040 0.78 ± 0.14 0.92 ± 0.23 0.027 ± 0.0065*** 0.048 ± 0.011*** CXCL1 0.0031 ± 0.00067 1.5 ± 0.32 0.64 ± 0.16 0.25 ± 0.054*** 0.15 ± 0.024*** CXCL2 0.0079 ± 0.0025 0.61 ± 0.15 0.63 ± 0.16 0.18 ± 0.045*** 0.18 ± 0.038*** 6h CXCL3 0.0053 ± 0.0014 0.91 ± 0.12 0.82 ± 0.096 0.078 ± 0.017*** 0.067 ± 0.016*** IL6 0.0029 ± 0.00098 3.9 ± 0.80 3.9 ± 0.76 0.042 ± 0.017*** 0.060 ± 0.025*** IL8 0.0035 ± 0.0016 1.4 ± 0.29 1.2 ± 0.26 0.13 ± 0.027*** 0.15 ± 0.026*** ISG20 0.44 ± 0.098 5.1 ± 1.3 6.4 ± 2.3 0.67 ± 0.12*** 0.80 ± 0.12*** OLR1 0.019 ± 0.0044 3.1 ± 0.48 3.3 ± 0.63 0.37 ± 0.075*** 0.23 ± 0.059*** PTGS2 0.053 ± 0.0089 1.1 ± 0.16 0.96 ± 0.12 0.069 ± 0.0087*** 0.055 ± 0.0074***

78 A time (h) 1 2 6 B DUSP1 siRNA 2 + + + + + + time (h) 1 2 6 DUSP1 siRNA 1 + + + + + + DUSP1 siRNA 2 + + + LMNA siRNA + + + + + + DUSP1 siRNA 1 + + + Dex + + + + + + + + + LMNA siRNA + + + IL1B - + + + + + + - + + + + + + - + + + + + + IL1B+Dex + + + + + + + + + 300 *** 100 *** 200 ** ** ** ** * *** *** * 100 50

CCL2 ** *** *** *** CCL2 0 *** 0 200 *** 100 ** ** *** * 100 * ** 50 **

CCL20

CCL20 0 *** 0 * * 200 ** *** *** 100 ** 100 50

CSF2 *** * CSF2 0 *** 0 150 *** *** *** *** *** *** *** 100 100 ** * 50 * ** 50

CXCL1 *** CXCL1

siRNA) 0 0 200 ** *** 150 * ** *** ** *

*** 100 on gene/GAPDHon

LMNA 100 * *** * relevant siRNA)

 50

CXCL2

CXCL2

***  Dex

B B 0 0

B B

Gene/GAPDH 1

* - IL1 200 *** *** *** *** IL 100 *** *** *** *** (% 100 * * ** (% 50

CXCL3 *** CXCL3 ** 0 *** Effectof 0 300 *** ** *** 100 200 ** ** **

IL6

** IL6 100 50 ** ** * 0 *** *** 0 ** 200 ** *** * 100 * *** ** *** **

IL8 100 IL8 50 *** *** * 0 *** 0 *** * *** * *** 200 *** *** 100 * ** *** *** * 100 ** *** *** ** * 50 ** PTGS2 ** *** PTGS2 0 *** 0

C time (h) 1 2 6 D DUSP1 siRNA 2 + + + + + + time (h) 1 2 6 DUSP1 siRNA 1 + + + + + + DUSP1 siRNA 2 + + + LMNA siRNA + + + + + + DUSP1 siRNA 1 + + + Dex + + + + + + + + + LMNA siRNA + + + IL1B - + + + + + + - + + + + + + - + + + + + + IL1B+Dex + + + + + + + + +

200 ** ** 100

* **   100 * ** 50

ISG20

ISG20

* B

1 -

1B 0

siRNA) 0 -

*** IL IL 300 *** *** *** *** 100 200 *** (%

(% ***

* * gene/GAPDH

LMNA 50 **

relevant relevant siRNA) Gene/GAPDH

OLR1

OLR1 100 * EffectDexonof * 0 *** *** 0

Figure 3.10 Effect of DUSP1 targeting siRNAs on IL1B-induced inflammatory gene mRNA expression.

79 Figure 3.10 continued. A & C, A549 cells were incubated with LMNA (control) or DUSP1- specific siRNAs for 24 h before being treated with IL1B (1 ng/ml) or IL1B plus dexamethasone (1 μM) (Dex) as indicated. Cells were harvested after 1, 2 or 6 h for real-time PCR analysis of the indicated genes and GAPDH. Data (N = 5 - 13) normalized to GAPDH, were expressed as a percentage of LMNA siRNA plus IL1B-stimulated for each time and plotted as means ± S.E. Significance was tested using ANOVA with a Newman-Keul multiple comparison test. Significance between: LMNA control siRNA plus IL1B and each of the DUSP1 targeting siRNAs plus IL1B, and the LMNA control plus IL1B plus Dex is shown. Other comparisons are specifically indicated. *, p < 0.05; **, p < 0.01; ***, p < 0.001. ‘Early-phase’ genes are shown in A and ‘delayed response’ genes in C. B & D, For each inflammatory gene at each time, the effect of IL1B plus Dex expressed as a percentage of IL1B for each of the three individual siRNAs is plotted as a mean ± S.E. The percent IL1B plus dexamethasone/IL1B for the LMNA siRNA is compared with that for each of the DUSP1-specific siRNAs using ANOVA with a Dunnett's post test. *, p < 0.05; **, p < 0.01; ***, p < 0.001. ‘Early-phase’ genes are shown in B and ‘delayed response’ genes in D.

Thus, with the exception of OLR1, IL1B plus dexamethasone-induced DUSP1 expression exerts negative regulation of these inflammatory genes. However, this effect had completely waned by 6 h and none of the genes showed any significant effects of DUSP1 silencing when compared to

IL1B plus dexamethasone (plus LMNA siRNA) (Fig. 3.10A & C). At 2 h, the effect of DUSP1 knock-down was rather variable and intermediate with the effects at 1 and 6 h.

In order to correctly analyse the role of DUSP1 in the dexamethasone-dependent repression of inflammatory genes, it is necessary to consider the feedback control exerted by IL1B-induced

DUSP1. For example, DUSP1 exerts significant feedback inhibition of IL1B-induced CCL2 mRNA at 1 h post-treatment (Fig. 3.10A). At this time, IL1B-induced CCL2 expression was significantly repressed by dexamethasone and this was modestly reversed by the DUSP1 siRNAs

(Fig. 3.10A).

However, the repressive effect of dexamethasone was not changed when IL1B plus dexamethasone-induced expression of CCL2 was expressed as a percentage of the IL1B-induced

CCL2 for the LMNA and DUSP1 targeting siRNAs (Fig. 3.10B, top left panel). This effect was

80 more evident in respect of CXCL3, IL6 and IL8 at 1 h. In each case, dexamethasone produced a significant repression and the DUSP1-targeting siRNAs increased this mRNA expression relative to IL1B plus dexamethasone (plus LMNA siRNA)-treated cells (Fig. 3.10A). However, when the effect of dexamethasone in the presence of each siRNA (LMNA vs DUSP1 siRNA1 and DUSP1 siRNA2) was compared, there was no difference in the overall repression (Fig. 10B). Thus, despite a clear feedback inhibitory role in the mRNA expression of these genes, DUSP1 had no additional role in the repression of CCL2, CXCL3, IL6 and IL8 mRNAs by dexamethasone at 1 h. A similar result was observed for OLR1 (Fig. 3.10D).

However, the effect of DUSP1 silencing was contrasting in respect of CXCL1, CXCL2, PTGS2 and CSF2. For each of these genes, IL1B-induced gene expression at 1 h was significantly attenuated by dexamethasone, and in each case, DUSP1 knock-down enhanced their mRNA expression (Fig. 3.10A). Thus, DUSP1 exerts a negative regulatory effect on these mRNAs in the presence of IL1B plus dexamethasone. The relative repressive effect of dexamethasone was next considered in the context of each siRNA treatment (Fig. 3.10B). For CXCL1 and CXCL2, the level of repression produced by dexamethasone plus LMNA siRNA was significantly higher than with the two DUSP1 targeting siRNAs (Fig. 3.10B). Indeed, in the presence of the DUSP1 targeting siRNAs dexamethasone did not produce any repression. A similar trend was observed for CSF2.

Thus, the repressive effect of dexamethasone on CXCL1 and CXCL2, and possibly CSF2, at 1 h, is mostly dependent on DUSP1. Similarly, ISG20, was also repressed by dexamethasone at 1 h and in the presence of the DUSP1-targeting siRNAs there was no repression (Fig. 3.10C & D). In respect of PTGS2 mRNA, dexamethasone-induced repression was significantly reversed by the two DUSP1-targeting siRNA at 1 h (Fig. 3.10A & B). However, even in the presence of DUSP1

81 siRNAs, there was still significant repression by dexamethasone (Fig. 3.10A & B). Thus, DUSP1 does not account for the full repressive effect of dexamethasone on PTGS2 mRNA at 1 h.

All 9 primary response genes and the two delayed response genes induced by IL1B were profoundly repressed by dexamethasone at 6 h (Fig. 3.10A & C). However, in all cases DUSP1 knock-down did not produce any effect. This contrasts with 2 h, where intermediate effects, between 1 and 6 h, were observed (Fig. 3.10A). It is important to note that the percent repressive effect of dexamethasone for the 9 primary response genes at 6 h was apparently reduced with

DUSP1 knock-down relative to the repression achieved in the presence of LMNA siRNA (Fig.

3.10B). However, this effect was mainly attributed to the lower expression of each mRNA that is produced following DUSP1 knock-down in the presence of IL1B (Fig. 3.10A). As a consequence, these data do not support a role for DUSP1 in the dexamethasone-dependent repression of the 9 primary response genes at 6 h. Conversely, while IL1B plus dexamethasone-induced ISG20 mRNA expression was unaffected by DUSP1 knock-down (Fig. 3.10C & D), the dexamethasone- dependent repression of OLR1 was modestly reduced by the DUSP1 targeting siRNA (Fig. 3.10D).

3.3.7 Effect of DUSP1 knock-down on IL1B-induced protein expression in the absence and presence of dexamethasone.

Supernatants from the experiments in Figs. 3.8 - 3.10 were analyzed for cytokine release. The release of CSF2, CXCL1, IL6 and IL8 was significantly enhanced following IL1B treatment at 6 h and this was significantly repressed by dexamethasone in a manner that was unaffected by the

LMNA control siRNA (Table 3.4). However, because of the need to use sub-confluent cells to achieve optimal siRNA transfection, the release of CSF2 at 2 h was below the detection limit of the assay and hence could not be analysed (Fig. 3.11A). IL1B-induced release of CSF2 and CXCL1 was not affected by the two DUSP1-targeting siRNAs, whereas the release of IL6 and IL8 showed

82 a trend toward enhanced expression (Fig. 3.11), and reached significance in respect of IL6 at 6 h with one DUSP1 siRNA. In the case of PTGS2 protein, while LMNA siRNA produced no effect,

IL1B-induced PTGS2 expression was enhanced by DUSP1 targeting siRNAs at 2 h (Fig. 3.11C).

However, at 6 h, IL1B-induced PTGS2 expression was unaffected by DUSP1 knock-down.

Table 3.4 Effect of LMNA targeting siRNA (LsiRNA) on IL1B-induced inflammatory protein expression. A549 cells were incubated with LMNA-specific siRNA for 24 h before being treated with IL1B (1 ng/ml) or IL1B plus dexamethasone (1 μM) (Dex) as indicated. Cell supernatants were also harvested after 2 or 6 h for measurement of cytokine/chemokine release. Data (N = 11 - 13) expressed in pg/ml and presented as means ± S.E. Note: release of CSF2 at 2 h was below the detection limit of the assay and hence not analysed further. Significance, relative to time-matched IL1B and IL1B plus LMNA siRNA-treated samples, was tested by ANOVA with a Bonferroni post test *, p < 0.05; **, p < 0.01; ***, p < 0.001.

Protein Release (pg/ml) Time Gene Symbol NS IL1B IL1B + LSiRNA IL1B + Dex IL1B + Dex + LsiRNA CXCL1 16 ± 3.9 490 ± 130 460 ± 120 270 ± 80*** 280 ± 77*** 2h IL6 0.27 ± 0.13 4.0 ± 0.94 3.7 ± 0.88 0.90 ± 0.34*** 0.78 ± 0.20*** IL8 17 ± 5.3 430 ± 130 410 ± 120 150 ± 37** 170 ± 43* CSF2 0.96 ± 0.27 21 ± 3.03 19 ± 2.6 3.1 ± 0.35*** 3.1 ± 0.34*** CXCL1 100 ± 23 9400 ± 1600 8900 ± 1600 2500 ± 740*** 2400 ± 660*** 6h IL6 0.38 ± 0.25 68 ± 13 48 ± 7.1 1.5 ± 0.79*** 2.4 ± 1.0*** IL8 130 ± 43 5800 ± 1000 4800 ± 920 770 ± 170*** 840 ± 170***

In terms of repression by dexamethasone, the release of each cytokine/chemokine was significantly repressed by dexamethasone. However, DUSP1-targeting siRNAs generally increased the release of each cytokine/chemokine in the presence of dexamethasone, and this was significant for CXCL1 at 2 and 6 h and IL8 at 2 h (Fig. 3.11A). When expressed as a percentage of IL1B for each siRNA

(Fig. 3.11B), the repression produced by dexamethasone at 6 h was significantly attenuated by both DUSP1 targeting siRNAs for CSF2, CXCL1 and IL8, and by one DUSP1-targeting siRNA for IL6.

83 A time (h) 2 6 B DUSP1 siRNA 2 + + + + time (h) 2 6 DUSP1 siRNA 1 + + + + DUSP1 siRNA 2 + + LMNA siRNA + + + + DUSP1 siRNA 1 + + Dex + + + + + + LMNA siRNA + + IL1B - + + + + + + - + + + + + + IL1B+Dex + + + + + + 40 *** *** 100 20 50 ***

CSF2 * CSF2 0 *** 0 1500 *** 15000 ** 150 *** 1000 ** * *** * 10000 *** 100 * ** *** 500 5000 50

CXCL1 *** CXCL1 0 0 0

12 ** 150 *** 150

+relevant siRNA (pg/ml)

Release 8 *** 100 ** *** 100 EffectDex of

IL6

4 50 IL6 50 * IL1B IL1B 0 ** 0 *** 0 1000 *** *** (% 150 *** 10000 ** *** 100

IL8 IL8 500 *** *** 5000 50 ** 0 0 *** 0

C time (h) 2 6 time (h) 2 6 DUSP1 siRNA 2 + + + + LMNA siRNA + + + + DUSP1 siRNA 1 + + + + Dex + + + + LMNA siRNA + + + + IL1B - + + + + - + + + + Dex + + + + + + IL1B - + + + + + + - + + + + + + PTGS2 PTGS2 GAPDH GAPDH

Figure 3.11 Effect of DUSP1 targeting siRNA on IL1B-induced inflammatory gene protein release/expression. A, A549 cells were incubated with LMNA (control) or DUSP1-specific siRNAs for 24 h before being treated with IL1B (1 ng/ml) or IL1B plus dexamethasone (1 μM) (Dex) as indicated. Supernatants were harvested after 2 or 6 h for cytokine/chemokine release measurement. Data (N = 9 - 15) expressed as pg/ml are plotted as means ± S.E. Note: release of CSF2 at 2 h was below the detection limit of the assay and hence not shown. Significance was tested between the LMNA control siRNA plus IL1B and each of the DUSP1 targeting siRNAs plus IL1B, and LMNA control plus IL1B plus Dex using ANOVA with a Newman-Keul multiple comparison test. Other comparisons are specifically indicated. *, p < 0.05; **, p < 0.01; ***, p < 0.001. B, For each cytokine/chemokine at each time, the effect of IL1B plus Dex was expressed as a percentage of IL1B for each of the three individual siRNAs and is plotted as a mean ± S.E. The percent IL1B plus dexamethasone/IL1B for the LMNA siRNA is compared with that for each DUSP1-specific siRNAs using ANOVA with a Dunnett's post test. *, p < 0.05; **, p < 0.01; ***, p < 0.001. C, Cells were treated as in A and harvested for western blot analysis of PTGS2 and GAPDH. Blots representative of at least 4 such experiments are shown.

Thus, a role for DUSP1 in the dexamethasone-dependent repression of CSF2, CXCL1, IL6 and

IL8 protein release was indicated. However, because this reversal of repression was very partial at

84 6 h, the existence of non DUSP1-dependent mechanisms of repression is predicted (Fig. 3.11B).

Conversely, the reversal of repression produced by the DUSP1 siRNAs at 2 h was complete for

CXCL1 and near complete for IL8 at 2 h (Fig. 3.11A & B). Thus, DUSP1 provides all, or the majority, of the repressive effect of dexamethasone on CXCL1 and IL8 at 2 h, but may have a lesser role at 6 h. The effect on PTGS2 protein was similar to that for IL6 release. IL1B plus dexamethasone-induced PTGS2 expression was not affected by LMNA siRNA (Fig. 3.11C).

However, despite a clear feedback role, DUSP1 knock-down had no effect on dexamethasone- dependent repression of PTGS2 protein at any time (Fig. 3.11C).

3.4 Discussion

In this chapter the feedback inhibitory and repressive role of IL1B and IL1B plus dexamethasone- induced DUSP1 was investigated on inflammatory gene expression. The mRNA expression of 11

IL1B-induced inflammatory genes, whose repression by dexamethasone was sensitive to translational blockade, was only modestly repressed by p38, MEK1/2 and JNK inhibitors when used alone. However, combined inhibition of all the three MAPKs together produced almost complete inhibition of the mRNA and, where tested, protein expression of all 11 genes. This observed difference between the effects of individual and combined MAPK inhibition could be attributed to the pathway redundancy in the induction of these 11 IL1B-induced inflammatory genes. Alternatively, cross-regulatory control between MAPK pathways may diminish the apparent role of any single MAPK pathway following its inhibition (362, 363). Thus, simultaneous inhibition of all three MAPK pathways may overcome this issue and also represents a highly effective means of inhibiting inflammatory gene expression, one that is also adopted by glucocorticoids. Likewise, MAPK activation and inflammatory gene expression were significantly inhibited by DUSP1 over-expression. Since, expression of the 11 IL1B-induced inflammatory

85 genes was inhibited by MAPK inhibitors and DUSP1 over-expression, the current data suggest that the expression of these 11 IL1B-induced inflammatory genes is MAPK-dependent. Thus, these genes should be valid targets to explore the function of DUSP1 that is induced by both IL1B and dexamethasone.

To understand the mechanism(s) of glucocorticoid-mediated inflammatory gene repression by

DUSP1, it is essential to first consider the feedback regulatory role of DUSP1 (362, 363).

Following IL1B treatment, loss of DUSP1 enhanced the phosphorylation of MAPK at 1 h.

However, there was no obvious effect at 2 or 6 h due to the loss of DUSP1. These data therefore suggest that IL1B-induced ERK, p38 and JNK MAPKs are only transiently regulated by DUSP1 via negative feedback processes. Similarly, siRNA-mediated loss of DUSP1 also enhanced the expression of 9 primary response genes shown in Fig. 3.2 A, plus ISG20 and OLR1, at 1 h and this was significant for at least 7 genes. This was in line with previous observations indicating the feedback regulatory role of DUSP1 in the regulation of multiple inflammatory genes, including

CSF2, IL6, IL8 and PTGS2 in A549 cells (284, 304, 353). However, this enhancement of inflammatory mRNA by DUSP1 loss was not seen at, or after, 2 h of IL1B treatment. In contrast,

IL1B-induced expression of many of the primary response genes, but not the two delayed response genes, was significantly reduced relative to IL1B plus LMNA treated cells by DUSP1 knock-down at 6 h post-IL1B treatment. These data therefore point to the possibility that enhanced phosphorylation of MAPKs (at 1 h) in the presence of DUSP1 siRNA may further enhance negative feed-forward regulation, for example, the expression of an inhibitory protein (possibly at

2 or 4 h), which may consequently produce enhanced loss of IL1B-induced inflammatory mRNAs

(at 6 h) in DUSP1 inhibited cells. Indeed, additional layers of MAPK-dependent regulatory control

86 exist in the form of ZFP36, which binds to AUUUA-containing mRNAs, e.g. CSF2, IL8 or PTGS2, to promote mRNA destabilization and translational silencing (364, 365). ZFP36 is strongly and transiently induced by pro-inflammatory stimuli (365), including IL1B, in A549 cells, where the expression is MAPK-dependent (366). Furthermore, enhanced MAPK activation following

DUSP1 loss is associated with enhanced ZFP36 expression (320). This in turn may lead to enhanced feed-forward control of ARE-containing inflammatory genes by ZFP36. Since the

3’UTR of all 9 of the primary response genes shown in Fig. 3.2A have one or more AUUUA motifs, which would be expected to bind ZFP36, it may explain the observed attenuation of inflammatory mRNA expression at 6 h post-IL1B treatment in DUSP1 inhibited cells. These data also offer a mechanistic explanation for the effect of glucocorticoids on ZFP36 expression. While

ZFP36 is induced by pro-inflammatory stimuli, such as IL1B (366), glucocorticoids, by reducing

MAPK activity, modestly reduce inflammatory stimuli-induced expression of ZFP36 (366, 367).

This glucocorticoid-induced attenuation of ZFP36 may seem detrimental in the context of anti- inflammatory action of glucocorticoids. However, ZFP36 expression is also modestly induced by glucocorticoids (366, 368, 369). Thus, by balancing the potential losses in ZFP36 expression due to the repression of MAPKs, glucocorticoids maintain the expression, or activity, of key effectors involved in regulatory control. Without the maintenance of such regulatory circuits, glucocorticoids could enhance inflammatory responses.

In terms of the repression of MAPKs by dexamethasone, DUSP1 plays a non-redundant role as the major effector in the early onset of repression of all three MAPK pathways following IL1B plus dexamethasone co-treatment. This was correlated with enhanced expression of DUSP1 in the presence of IL1B plus dexamethasone. However, the role of DUSP1 in MAPK repression was

87 diminished by 2 h, and, at 6 h, there was no obvious effect of DUSP1 knock-down. This suggests that there must be additional DUSP1-independent mechanisms by which dexamethasone exerts repression of MAPKs. In this regard, glucocorticoid-induced effector genes, TSC22D3 and

CDKN1C, which may inhibit Ras-Raf activation and JNK signalling, respectively, are both induced by dexamethasone in A549 cells and may play an important role in dexamethasone- induced repression of MAPKs (282, 350, 363, 370-373). In addition, dexamethasone-mediated enhancement of Dok-1, SLAP and Dexras1 inhibit the activation of signalling transduction cascades through a variety of different mechanisms and may contribute to MAPK repression by dexamethasone (275-277). Furthermore, the data presented here are also supported by a study showing similar results where MAPK phosphorylation was only partially affected by DUSP1 knock-down following dexamethasone pre-treatment (114).

In the case of inflammatory gene repression by dexamethasone, the repression of IL1B-induced

CXCL1 and CXCL2 mRNAs was almost completely prevented by DUSP1 knock-down at 1 h, with a similar trend for CSF2. In contrast, dexamethasone-induced repression of CCL2, CXCL3,

IL6, IL8 and OLR1 at 1 h was not affected by the loss of DUSP1. This may be due to insufficient knock-down of DUSP1. However, this may seems unlikely due to the fact that DUSP1-targeting siRNAs prevented the dexamethasone-dependent inhibition of MAPK phosphorylation and the repression of CXCL1, CXCL2 mRNA at 1 h. Thus, DUSP1-independent mechanisms may play a role in dexamethasone-induced repression of CXCL3, IL6, IL8 and OLR1. Moreover, even though

DUSP1 was involved in the feedback control of IL1B-induced PTGS2 at 1 h, dexamethasone- dependent repression was only moderately prevented in the presence of DUSP1 siRNAs, and thus, suggests a role for both DUSP1-dependent and DUSP1-independent mechanisms of repression by

88 dexamethasone. In respect of dexamethasone-dependent repression at 6 h, the repression of all the mRNAs was enhanced relative to earlier times. However, DUSP1 knock-down did not affect the repression of any of these genes and therefore these data do not support a role for DUSP1 in dexamethasone-induced repression occurring at 6 h. Thus, following the early phase of DUSP1- dependent repression, additional effector processes must be induced by dexamethasone, in addition to DUSP1, that either play a predominant role or act redundantly with DUSP1 to produce inflammatory gene repression. For example, while DUSP1 protein expression was induced within

30 min-1 h post-IL1B plus dexamethasone treatment, the expression of TSC22D3, which negatively regulates the activation of NF-κB, AP-1 and possibly MAPKs, is also induced within

1-2 h following IL1B plus dexamethasone treatment (370, 372). Thus, in addition to DUSP1,

TSC22D3 may also play a role in dexamethasone-dependent repression of inflammatory genes at and after 1 h.

For IL6, IL8 and PTGS2 protein expression, DUSP1 knock-down showed a trend towards increased expression and this was in line with the previously described feedback inhibitory role for DUSP1 in the regulation of these genes (284). At the level of CXCL1 protein release, dexamethasone-dependent repression was totally prevented by the loss of DUSP1 at 2 post-IL1B treatment. This confirms that DUSP1 is a predominant effector of repression at this time. However, by 6 h post-IL1B, dexamethasone-dependent repression of CXCL1 release was only partially reversed by the loss of DUSP1. Similarly, a role for DUSP1 in the dexamethasone-dependent repression of IL8 release at 2 h was clearly evident. By 6 h post-IL1B plus dexamethasone treatment, both IL8 and CSF2 release showed a very modest inhibition of repression following

DUSP1 knock-down. Thus, a role for DUSP1 in the repression by dexamethasone was clearly

89 confirmed for CXCL1 and IL8 release at 2 h, but only a modest role can be established at 6 h.

Conversely, dexamethasone-induced repression of IL6 and PTGS2 at both 2 and 6 h may be predominantly mediated by DUSP1-independent mechanisms of repression.

In conclusion, the data presented in this chapter demonstrate that, in A549 cells, glucocorticoid- induced DUSP1 plays a major role in the repression of MAPKs, selected inflammatory mRNAs and proteins at early times. However, this role is only transient, and by 6 h, DUSP1-independent mechanisms of repression may dominate with respect to the repression of MAPKs and inflammatory genes by dexamethasone. In summary, these current data, contrary to evolving dogma that DUSP1 is a major effector of glucocorticoid action, suggest that DUSP1 only has a transient, more minor, role compared to DUSP1-independent mechanisms of repression.

Therefore, these data emphasize the need to continue the characterization of glucocorticoid- induced effector responses that are important in the repression of inflammatory genes and to explore their roles relative to known effectors such as DUSP1. Only by understanding the detailed mechanism(s) of the anti-inflammatory response of glucocorticoids, the anti-inflammatory NR3C1 agonist therapies for the treatment of asthma can be improved.

90 Chapter 4 : Downregulation of the mitogen-activated protein kinase phosphatase, DUSP1 limits TNF expression through enhanced expression of tristetraprolin (TTP; ZFP36)

Data presented in this chapter have been published:

Shah S, Mostafa MM, McWhae A, Traves SL, Newton R. Incoherent feed-forward control of TNF by tristetraprolin (ZFP36) is limited by the mitogen-activated protein kinase phosphatase, DUSP1: Implications for regulation by glucocorticoids. J. Biol. Chem. 2016;291(1):110-25

Copyright  Journal of Biological Chemistry.

This work was reprinted with permission based on the J. Biol. Chem. copyright permission policy.

Mostafa MM contributed to Fig. 4.2, 4.5 and 4.6

McWhae A contributed to Fig. 4.2

Traves SL contributed to 4.5 and 4.6

91 4.1 Rationale

In the previous chapter, knock-down of IL1B-induced DUSP1 expression transiently enhanced the appearance of phosphorylated MAPKs and this was correlated with increased expression of inflammatory mRNAs at 1 h post-IL1B treatment. However, the expression of multiple inflammatory mRNAs at 6 h post-IL1B treatment was decreased by the loss of DUSP1. This observation, while initially unexpected, is consistent with the concept that MAPKs may increase

ZFP36 expression to subsequently down-regulate ARE-containing mRNAs. This is tested in the current chapter (Fig. 4.1). Indeed, in A549 cells, ZFP36 is strongly and transiently induced by

IL1B in a MAPK-dependent manner (366). Additionally, enhanced activation of MAPKs following DUSP1 loss is associated with augmented expression of ZFP36 (320). Thus, enhanced

ZFP36 expression may further induce the feed-forward control of the 9 primary response ARE- containing inflammatory mRNAs (CCL2, CCL20, CSF2, CXCL1, CXCL2, CXCL3, IL6, IL8 and

PTGS2) (shown in the previous chapter). ZFP36-mediated feed-forward control, in DUSP1 inhibited cells, may consequently produce enhanced loss of primary response inflammatory mRNAs at 6 h post-IL1B treatment. In this context, post-transcriptional regulation of the pro- inflammatory cytokine, TNF, is conferred via multiple AREs located in the 3’-UTR of the TNF mRNA (214, 374). This region is critical for regulating message stability and is targeted by several

RNA binding proteins, including ZFP36 (219). ZFP36 negatively controls TNF expression by promoting mRNA deadenylation and degradation with consequent reductions in TNF biosynthesis

(375-377). Thus, TNF was used as a model, ZFP36 target gene, to explore the relationship(s) between the regulation of MAPK activation by DUSP1 with the expression of ZFP36 and the subsequent effects on downstream gene expression in A549 cells

92 IL1B treatment IL1B

DUSP1 MAPK Feed-forward control by ZFP36 TNF gene expression

Figure 4.1 Enhanced inflammatory gene expression by IL1B: Feedback control by DUSP1 and feed-forward control by ZFP36. IL1B treatment results in the activation of MAPKs. This, along with the activation of other signalling pathway and inflammatory transcription factors, e.g. NF-κB and AP-1 (not shown), enhances the expression of inflammatory genes (e.g. TNF), as well as the negative feedback regulator, DUSP1, and the feed-forward regulator, ZFP36. By binding and promoting mRNA decay, and/or translational arrest, of ARE-containing mRNAs, ZPF36 is a negative regulator of inflammatory gene expression. Following MAPK activation, the increased expression of DUSP1 is one mechanism by which MAPK activity is restored to basal. Expression of ZFP36 depends on MAPK activation and this limits the expression of ARE-containing inflammatory genes, such as TNF.

4.2 Hypothesis

Since, the loss of DUSP1 in chapter three was associated with increased loss of IL1B-induced inflammatory mRNAs at 6 h, the hypothesis for this chapter is that siRNA-mediated loss of

DUSP1, by augmenting MAPK phosphorylation at 1 h, increases ZFP36 expression, leading to the enhanced loss of ARE-containing mRNAs at 6 h. The implications of these programme are also investigated in the context of glucocorticoid treatment.

4.3 Results

4.3.1 Characterization of TNF expression in the presence of IL1B and dexamethasone

TNF mRNA was rapidly induced by IL1B (Fig. 4.2A), reached a peak at 2 h post-stimulation, and then declined sharply towards basal levels over the following 4 h. In addition, the expression of unspliced nuclear TNF RNA was also transiently augmented by IL1B at 1 and 2 h and suggests

93 that rapid enhancement of TNF transcription may contribute to the increase in TNF mRNA (Fig.

4.2B). To assess the role of post-transcriptional mechanisms in IL1B-induced TNF mRNA expression, the stability of TNF was assessed by the actinomycin D chase method. IL1B-induced

TNF mRNA was very stable at 30 min post-stimulation (Fig. 4.2C). However, treatment with IL1B for 90, 120 and 180 min, resulted in reduced TNF mRNA expression (a loss of ~50% of initial levels (t = 0)) within 30 - 40 min of actinomycin D treatment (Fig. 4.2C). Conversely, at short, 30 min, IL1B treatment, the mature (spliced) TNF mRNA levels continued to increase following actinomycin D treatment. This observed enhancement in TNF mRNA could be due to the fact that while actinomycin D inhibits RNA Pol II-dependent transcription, the production of TNF mRNA still continues for a very short period of time following actinomycin D treatment. This brief production of TNF mRNA could be mainly attributed to the accumulation of immature/unspliced

TNF RNA (Fig. 4.2B). This explanation is also in line with the fact that splicing and processing of immature mRNAs can take several minutes, or often longer, to produce mature mRNAs (378).

Thus, the actinomycin D chase experiments may be considered as reflecting post-transcription

RNA processing and maturation as well as mRNA degradation. Moving to TNF protein, TNF was not released into the supernatant under basal conditions, and only very low levels of TNF protein release were detected at 2 h following IL1B treatment (Fig. 4.2D, left panel). TNF release reached a peak around 4 h post-IL1B treatment and maintained at this level for up to 18 h. In order to analyse the expression of TNF protein inside or associated with the cells, following the removal of supernatants, cells were washed with PBS, and then lysed in a soft-lysis buffer (containing protease and phosphatase inhibitors) that was compatible with the ELISA. Following ELISA analysis of cell lysates, cell-associated TNF protein was first detected at 1 h post-IL1B treatment

(Fig. 4.2D, right panel). This was further enhanced at 2 h, before reaching a peak at 4 h, and then

94 declining towards basal levels by 6 h. Since, TNF is initially expressed as a membrane bound uncleaved form that then gets cleaved by the metalloproteinase, TNF-converting enzyme (TACE), the expression of uncleaved TNF was also examined by western blotting (Fig. 4.2E, left panel)

(379). Uncleaved TNF was detected very rapidly within 1 h post-IL1B treatment, reached a maximum level by 2 h, and returning to basal levels by 6 h (Fig. 4.2E, left panel).

IL1B-induced TNF mRNA was modestly repressed at all times by dexamethasone (Fig. 4.2A).

Since dexamethasone also produced a partial attenuation of unspliced nuclear TNF RNA, these data suggest that the repressive effect of dexamethasone on TNF mRNA is predominantly attributed to transcriptional repression (Fig. 4.2B). Equally, the t½ of TNF mRNA, following 90 min of IL1B treatment, was around 30 - 40 min, and this was further decreased by dexamethasone post-actinomycin D treatment (Fig. 4.2C, right panels). In the case of TNF protein, dexamethasone almost completely repressed IL1B-induced cell-associated, uncleaved and secreted TNF at all times (Fig. 4.2D & E). Thus, in addition to transcriptional and post-transcriptional repressive mechanisms, repression of TNF by dexamethasone may also involve translational, or possibly post-translational, mechanisms of repression.

95 A B 50 1.5 100 100 NT 1.0 IL1B * 25 50 * * * 50 IL1B+Dex RNA 0.5 *

mRNA *

IL1B+Dex each time) each

unTNF/U6

IL1B+Dex time) each

(% ofat(% IL1B (% ofat (% IL1B TNF/GAPDH 0 0 0.0 0 0 2 4 6 1 2 4 6 0 2 4 6 1 2 4 6 time (h) time (h) time (h) time (h) C 300 30 min 150 90 min IL1B time post Act D (min) 45 100 200 IL1B+Dex Dex + 50 150 IL1B + + 100 * 0 0 100 30 150 120 min 150 180 min

mRNA 50 15

mRNA

(% t =t 0) (%

mRNA (% t =t 0) (%

100 100 =t (%0)

TNF/GAPDH TNF/GAPDH 0 * * * TNF/GAPDH 0 50 50 0 20 40 60 0 0 time (min) 0 15 30 45 0 15 30 45 time (min) time (min) D E time (h) 2 time (h) 1 1 2 6 Dex + 200 IL1B - + + + IL1B - + + Supernatant 30 Cell associated Uncleaved TNF Uncleaved TNF 20

100 GAPDH GAPDH TNF

TNF 10

(pg/ml)

(pg/well) release release

***

** * *** *** *** ** * * *** * 0 1.5 100 0 ** 0 5 10 15 20 0 2 4 6 1.0 50

time (h) TNF

0.5 TNF

protein

protein

/GAPDH (% IL1B) (% NT IL1B IL1B+Dex /GAPDH 0.0 0 Figure 4.2 Characterization of TNF expression in the presence of IL1B and dexamethasone. A549 cells were either not treated (NT) or stimulated with IL1B (1 ng/ml) or a combination of IL1B and dexamethasone (Dex, 1 μM) as indicated. Cells were harvested at the indicated times for real-time PCR analysis of (A) TNF and GAPDH mRNA or (B) unspliced nuclear (un) TNF RNA and U6 RNA. Data (N = 4), normalized to GAPDH or U6, are plotted as means ± S.E. The effect of IL1B + dexamethasone for TNF mRNA (A right panel) (N = 12) and unspliced nuclear (un) TNF RNA (B right panel) (N = 4) is plotted as a percentage of IL1B for indicated times. Significance, relative to time matched IL1B-treated samples was tested using a paired t-test. *, p < 0.05. C, A549 cells were treated with IL1B (1 ng/ml) for 30, 90, 120 and 180 min. Actinomycin D (Act D, 10 μg/ml) was then added (t = 0) and the cells were harvested as indicated. RNA was extracted for real-time PCR analysis of TNF and GAPDH. Data (N = 3) were normalized to GAPDH and are plotted as a percentage of t = 0 for each treatment as means ± S.E. (left panels). A549 cells were also treated with IL1B (1 ng/ml) or IL1B plus dexamethasone (Dex, 1 μM) for 90 min. After 90 min, actinomycin D (Act D 10 μg/ml) was added (t = 0) and the cells were harvested at the indicated times for real-time PCR analysis of TNF and GAPDH. Data (n = 4) were normalized to GAPDH and are plotted as a percentage of t = 0 for each treatment as means ± S.E. (middle panel). The effect of IL1B plus dexamethasone at 45 min post Act D treatment is plotted as a percentage of t = 0 for 45 min. Significance, relative to IL1B-treated samples was tested using a paired t-test. *, p < 0.05 (right panel). D, Supernatants (1 ml), from the cells in A, were harvested for TNF release measurement (left panel). Alternatively, cells were harvested after 1, 2, 4 or 6 h in soft lysis buffer and total protein was prepared for the detection of cell-associated TNF via ELISA (right panel). Data (N = 4), expressed in pg/ml (for supernatant) and pg/well (cont.…)

96 Figure 4.2 continued. (for cell associated TNF), are plotted as means ± S.E. Significance, relative to time matched IL1B-treatment, was tested using a paired t-test. *, p < 0.05; **, p< 0.01; ***, p< 0.001. E, A549 cells were either not treated or stimulated with IL1B (1 ng/ml) for 1, 2 or 6 h (left panel) or with IL1B (1 ng/ml) or a combination of IL1B and dexamethasone (Dex, 1 μM) for 2 h (right panel) and total protein was prepared for western blot analysis of uncleaved (~25 kDa) TNF and GAPDH. Blots representative of 4 such experiments are shown. Following densitometric analysis, data (N = 4) were normalized to GAPDH and plotted as means ± S.E. Significance, relative to non-treated (left panel) or IL1B-treated (right panel) samples was tested using ANOVA with a Dunnett's post test (left panel) or a paired t-test (right panel) is indicated.*, p < 0.05;**, p< 0.01; ***, p< 0.001.

4.3.2 Analysis of DUSP1 expression in the presence of IL1B and dexamethasone

Following treatment of cells with IL1B in the absence and presence of dexamethasone, DUSP1 protein expression was analysed by western blotting. IL1B-induced DUSP1 reached a maximum level within 1 h, then declined steeply by 2 h, reaching basal levels by 6 h post-treatment (Fig.

4.3A). Conversely, dexamethasone alone produced a modest enhancement in DUSP1 protein at 1 h, and this was significantly enhanced by 2 h. In the case of IL1B plus dexamethasone co- treatment, DUSP1 protein was robustly induced at 1 h, and declined by 2 h post-treatment.

However, at 6 h, while IL1B did not induce any DUSP1 protein expression, dexamethasone alone significantly enhanced DUSP1 expression and this was maintained in the presence of IL1B. These results are in line with the data shown in the previous chapter.

97 A B time (h) ½ 1 2 6 time (h) 1 2 6 Dex + + + + + + + + Dex + + + + + + IL1B - - + + - - + + - - + + - - + + - - + + - - + + - - + + NS IL1B DUSP1 NS ZFP36 GAPDH 3 GAPDH *** 3.0 * * *** 2 * *** ** ** *** 1.5

protein 1

DUSP1 *** ***

ZFP36 protein /GAPDH *** *** * ** /GAPDH 0 0.0 Figure 4.3 Analysis of DUSP1 and ZFP36 expression in the presence of IL1B and dexamethasone. A & B, A549 cells were either not treated or stimulated with IL1B (1 ng/ml), dexamethasone (Dex, 1 μM) or a combination of the two. Cells were harvested at the times indicated prior to western blot analysis of DUSP1, ZFP36 and GAPDH. Blots representative of at least 4 such experiments are shown. Following densitometric analysis, data (N = 4) were normalized to GAPDH and plotted as means ± S.E. For A, Significance, using ANOVA with a Bonferroni's multiple comparison test is indicated. For B, Significance, using ANOVA with a Dunnett's post test is indicated. *, p < 0.05; **, p< 0.01; ***, p< 0.001. NS = nonspecific band.

4.3.3 Analysis of ZFP36 expression in the presence of IL1B and dexamethasone

ZFP36 was undetected in untreated cells, but the expression, comprised of a protein doublet, was significantly induced within 1 h of IL1B treatment (Fig. 4.3B). IL1B-induced ZFP36 expression was maintained at 2 h, and detected as a slowly migrating protein doublet, presumably due to the phosphorylation of ZFP36 (310, 380). The expression of IL1B-induced ZFP36 was considerably reduced by 6 h. Conversely, ZFP36 protein expression was not induced by dexamethasone at either

1 or 2 h (Fig. 4.3B). In fact, co-treatment with IL1B, dexamethasone significantly attenuated IL1B- induced ZFP36 expression at 1 and 2 h (Fig. 4.3B). However, at 6 h, ZFP36 was modestly induced by dexamethasone, and at this time there was no repression of IL1B-induced ZFP36 expression by dexamethasone (Fig. 4.3B).

4.3.4 Effect of MAPK inhibitors on DUSP1 and ZFP36 expression

A549 cells were treated with maximally effective concentrations of SB203580 (10 µM), U0126

(10 µM) or JNK inhibitor 8 (JNK-IN-8) (10 µM), which inhibit the p38, MEK1/2 (to inhibit ERK)

98 and JNK MAPK pathways respectively. SB203580 and JNK-IN-8 did not produce any clear effect on IL1B-induced DUSP1 expression at 1 h (Fig. 4.4A). However, UO126 produced a partial inhibition of DUSP1. In contrast, IL1B-induced DUSP1 protein expression was almost completely prevented in the presence of a combination of all three MAPK inhibitors, suggesting that IL1B- induced DUSP1 expression is MAPK-dependent.

In the case of IL1B-induced ZFP36 expression, while SB203580 produced a complete loss of

ZFP36 expression, U0126 and JNK-IN-8 both produced a modest, non-significant inhibition of

ZFP36 expression at each time (Fig. 4.4B). IL1B-induced ZFP36 expression was completely attenuated by the combined MAPK inhibitors at all times (Fig. 4.4B). Since IL1B-induced ZFP36 expression was MAPK dependent, these data suggest that reduced activity of MAPKs, by reducing the expression of ZFP36, could enhance the stability and expression of ARE-containing mRNAs, such as TNF (Fig. 4.1).

99 A B time (h) ½ 1 2 6 time (h) 1 2 6 Dex + + + + + + + + Dex + + + + + + IL1B - - + + - - + + - - + + - - + + IL1B - - + + - - + + - - + + DUSP1 NS NS ZFP36 GAPDH 3 GAPDH *** 3.0 * * *** 2 * *** ** ** *** 1.5

protein 1

DUSP1 *** ***

ZFP36 protein /GAPDH *** *** * ** /GAPDH 0 0.0

A B time (h) 1 2 6 JNK-IN-8 + + + + + + UO126 + + + + + + time (h) 1 SB203580 + + + + + + IL1B - + + + + + - + + + + + - + + + + + JNK-IN-8 + + NS UO126 + + ZFP36 SB203580 + + IL1B - + + + + + GAPDH NS DUSP1 ** 0.8 * ** GAPDH **

0.4

ZFP36

protein /GAPDH 0.0 Figure 4.4 Analysis of DUSP1 and ZFP36 expression in the presence of IL1B and MAPK inhibitors. A & B, A549 cells were either not treated, treated with IL1B (1 ng/ml) or pre-treated with either UO126, SB203580, JNK inhibitor 8 (JNK-IN-8) or a combination of UO126, SB203580 plus JNK-IN-8 each at 10 µM for 30 min prior to IL1B stimulation. Cells were harvested at the times indicated prior to western blot analysis of DUSP1 or ZFP36 and GAPDH. Blots representative of at least 4 such experiments are shown. Following densitometric analysis, data (N = 4) were normalized to GAPDH and plotted as means ± S.E. Significance, using ANOVA with a Dunnett's post test is indicated. *, p < 0.05; **, p< 0.01. NS = nonspecific band.

4.3.5 Analysis of DUSP1 expression in the presence of IL1B and dexamethasone in primary HBE cells

In order to further validate the findings obtained in A549 cells, the mRNA and protein expression of DUSP1 was examined in primary HBE cells. DUSP1 mRNA was robustly induced by IL1B and dexamethasone at 1 h, following which the mRNA expression declined rapidly at 2 h post- treatment (Fig. 4.5A, left panel). Whereas IL1B-induced DUSP1 mRNA had returned to basal levels at 18 h, DUSP1 mRNA was strongly induced by dexamethasone at 18 h (Fig. 4.5A, left panel). Thus, at 18 h, the mRNA expression of DUSP1 was mostly driven by dexamethasone, with no effect of IL1B. Equally, DUSP1 protein was also strongly induced by IL1B and IL1B plus dexamethasone (Fig. 4.5B, top panels). In the case of dexamethasone treatment alone, DUSP1 protein was induced at 6 and 18 h post-treatment and there was a little or no further effect of IL1B on dexamethasone-induced DUSP1 expression (Fig. 4.5B, top panels). Conversely, the expression

100 of DUSP1 protein was rapidly induced by IL1B, reached a peak within 1 h post-stimulation, and then declined slowly towards basal levels over the following 18 h.

A

*** 5.0 * NT 3 *** ** ** *** *** ** Dex *** *** * 2 * ** ** ** 2.5 * ** IL1B

* mRNA

ZFP36 * mRNA ** IL1B+Dex

DUSP1 1 * /GAPDH

/GAPDH 0 0.0 ½ 1 2 6 18 ½ 1 2 6 18 time (h) time (h) B time (h) ½ 1 2 6 18 Dex + + + + + + + + + + IL1B - - + + - - + + - - + + - - + + - - + + DUSP1 GAPDH ZFP36 NS GAPDH Figure 4.5 Analysis of DUSP1 and ZFP36 expression in primary human bronchial epithelial cells. A, HBE cells were either not treated (NT) or stimulated with IL1B (1 ng/ml), dexamethasone (Dex, 1 μM) or a combination of the two as indicated. Cells were harvested at the indicated times for real-time PCR analysis of DUSP1, ZFP36 and GAPDH. Data (N = 4), normalized to GAPDH, are plotted as means ± S.E. Significance, using ANOVA with a Bonferroni's multiple comparison test is indicated. *, p < 0.05; **, p< 0.01; ***, p< 0.001. B, HBE cells were treated as in A and harvested at the indicated times for western blot analysis of DUSP1, ZFP36 and GAPDH. Blots representative of at least 4 such experiments are shown. NS = nonspecific band.

4.3.6 Analysis of ZFP36 expression in the presence of IL1B and dexamethasone in primary HBE cells

ZFP36 mRNA was strongly induced by IL1B in HBE cells, and reached a peak at 1 h post- stimulation, before declining rapidly to basal levels by 2 h (Fig. 4.5A, right panel). Dexamethasone alone did not change ZFP36 mRNA expression until 18 h and also had no effect on IL1B-induced

ZFP36 mRNA expression at this time (Fig. 4.5A, right panel). In the case of IL1B plus dexamethasone co-treatment, ZFP36 mRNA was strongly induced at 1 h and declined steeply over the following 18 h post-treatment (Fig. 4.5A, right panel). The early induction of ZFP36 mRNA was predominantly mediated by IL1B, whereas the late phase expression of ZFP36 mRNA at 18

101 h was mainly dexamethasone-dependent. ZFP36 protein expression was rapidly induced by IL1B at 1 and 2 h post-treatment, thereafter IL1B-induced ZFP36 protein declined rapidly over the following 18 h (Fig. 4.5B, bottom panel). However, dexamethasone-induced ZFP36 expression was variable at 18 h. Dexamethasone either alone or in the presence of IL1B, did not produce any obvious effect on ZFP36 protein expression (Fig. 4.5B, bottom panel).

4.3.7 Characterization of TNF expression in the presence of IL1B and dexamethasone in HBE cells

Similar to A549 cells, TNF mRNA was rapidly enhanced by IL1B in HBE cells (Fig. 4.6A, left panel), reached a peak at 2 h and then declining sharply over the following 4 h. Since the CT values for TNF mRNA expression by real-time PCR was same (22 - 24 cycles) in both A549 and HBE cells (Fig. 4.7), these data suggest that TNF mRNA expression is similar in both cell types. In the case of protein expression, there was no TNF release under basal conditions. Treatment with IL1B produced low levels of TNF protein at 18 h (Fig. 4.6B). These data are in line with previous findings in HBE cells, where treatment of HBE cells with virus produced a very low release of

TNF (381). In addition, the expression of uncleaved ~25 kDa TNF was also examined by western blotting (Fig. 4.6C). Uncleaved TNF was rapidly induced by IL1B and first detected within 1 h, reached a peak at 2 h and was only marginally detectable by 6 h. Since all the blotting and detection conditions, including exposure time, were identical with the blots for A549 cells, presented in Fig.

4.2E, it can be concluded that, like TNF mRNA, uncleaved ~25 kDa TNF protein expression is also similar in A549 and HBE cells. Conversely, dexamethasone produced a modest repression of

IL1B-induced TNF mRNA at all times (Fig. 4.6A right panel). Similarly, uncleaved and soluble

TNF protein expression was also profoundly repressed by dexamethasone at 2 and 18 h respectively (Fig. 4.6B & C, right panel).

102 A B time (h) 18 NT Dex Dex + + IL1B IL1B+Dex IL1B - - + + 0.9 200 0.6 4 100

* * TNF

mRNA 0.3 2

(pg/ml)

release release

IL1B+Dex each time) each

TNF/GAPDH 0.0 ofat(% IL1B 0 0 1 2 6 18 0 5 10 15 time (h) time (h) time (h)

C time (h) 2 time (h) 1 1 2 6 Dex + IL1B - + + + IL1B - + + Uncleaved TNF Uncleaved TNF GAPDH GAPDH

4 *** 100 ** **

2 50

TNF

TNF

protein

/GAPDH

protein (% IL1B) (% 0 /GAPDH 0 Figure 4.6 Analysis of TNF expression in primary human bronchial epithelial cells. A, HBE cells were either not treated (NT) or stimulated with IL1B (1 ng/ml), dexamethasone (Dex, 1 μM) or a combination of the two as indicated. Cells were harvested for real-time PCR analysis of TNF and GAPDH. Data (N = 4), normalized to GAPDH, are plotted as means ± S.E. (left panel). The effect of IL1B + dexamethasone for TNF mRNA (N = 4) is plotted as a percentage of IL1B for indicated times (right panel). Significance, relative to time matched IL1B-treated samples was tested using a paired t-test. *, p < 0.05. B, Supernatants, from the cells in A were harvested for TNF release measurement. C, HBE cells were either not treated or stimulated with IL1B (1 ng/ml) for 1, 2 or 6 h (left panel). Alternatively, cells were not treated or stimulated with IL1B (1 ng/ml) or a combination of IL1B and dexamethasone (Dex, 1 μM) for 2 h (right panel). Total protein was prepared for western blot analysis of uncleaved (~25kDa) TNF and GAPDH. Blots representative of at least 4 such experiments are shown. Following densitometric analysis, data (N = 4) were normalized to GAPDH and plotted as means ± S.E. Significance, relative to non-treated (left panel) samples was tested using ANOVA with a Dunnett's post test or IL1B-treated (right panel) samples was tested using a paired t-test.*, p < 0.05;**, p< 0.01; ***, p< 0.001.

103

IL1B treatedIL1B stimulatedNot IL1B treatedIL1B A stimulatedNot B 1.0 E+1 1.0 E+1 TNF A549 cells TNF HBE cells 1.0 1.0 30 1.0 E-1 1.0 E-1 1.0 E-2 1.0 E-2 20 1.0 E-3 1.0 E-3

1.0 E-4 10

∆Reaction ∆Reaction 1.0 E-4 (Ct) Cycle 1.0 E-5 1.0 E-5 0 1.0 E-6 1.0 E-6

2 6 10 14 18 22 26 30 34 38 42 2 6 10 14 18 22 26 30 34 38 42

HBE A549 A549 Cycle (Ct) Cycle (Ct) Figure 4.7 Comparison of CT values for TNF expression in A549 and HBE cells generated by RT-PCR. A, SYBR green RT-PCR for TNF mRNA was performed in A549 (left panel) and HBE cells (right panel) for non-stimulated and IL1B-treated cells in duplicates. In each case, the samples were analysed for TNF mRNA using primers specific for TNF. The RT-PCR conditions were similar for both A549 and HBE cells. Plots of ∆Reaction against CT are shown. The positions of reaction profiles corresponding to non-stimulated and IL1B-treated samples are indicated. B, CT values for TNF mRNA expression in both A549 and HBE cells, obtained from A, representative of at least 4 RT-PCR results, is plotted as means ± S.E. 4.3.8 Effect of ZFP36 siRNA on IL1B-induced TNF mRNA expression

The role of ZFP36 in the regulation of TNF mRNA expression was examined by using ZFP36- targeting siRNAs. Control siRNA targeted to LMNA had no effect on IL1B and IL1B plus dexamethasone-induced ZFP36 and TNF expression at any time (Fig. 4.8A & B).

A B time (h) 1 2 6 LMNA siRNA + + + + + + time (h) 1 2 6 Dex + + + + + + LMNA siRNA + + + + + + IL1B - + + + + - + + + + - + + + + Dex + + + + + + 150 *** IL1B - + + + + - + + + + - + + + + *** ** *** *** ** ZFP36 NS 100

mRNA 50 GAPDH (%IL1B) TNF/GAPDH 0 Figure 4.8 Effect of LMNA siRNA on ZFP36 and TNF expression. A, A549 cells were incubated with or without LMNA-specific siRNA. After 24 h, cells were treated with IL1B (1 ng/ml) or IL1B plus dexamethasone (1 μM) (Dex) as indicated. Cells were harvested at 1, 2 or 6 h for western blot analysis of ZFP36 and GAPDH. Blots representative of at least 4 such experiments are shown. B, Cells treated as in A and harvested for real-time PCR analysis of TNF and GAPDH. Data (N = 4), normalized to GAPDH, are plotted as means ± S.E. Significance, using ANOVA with a Bonferroni's multiple comparison test is indicated. **, p< 0.01; ***, p< 0.001. NS = nonspecific band.

104 IL1B-induced ZFP36 expression was significantly inhibited by the two independent ZFP36- targeting siRNAs at all times tested (Fig. 4.9A). In respect of IL1B-induced TNF mRNA, ZFP36 knock-down did not produce any effect at 1 and 6 h, but significantly increased TNF mRNA expression at 2 h (Fig. 4.9A middle panel). This result further confirms that TNF mRNA is negatively regulated by ZFP36. Furthermore, IL1B-induced unspliced nuclear TNF RNA was also unaffected by ZFP36 silencing, and is consistent with post-transcriptional role for ZFP36 in the regulation of TNF expression (Fig. 4.9A, lower panels), and was further explored using the actinomycin D chase methodology. Thus, A549 cells were treated with LMNA or ZFP36-targeting siRNAs followed by treatment with IL1B for 90 min prior to actinomycin D chase (Fig. 4.9B).

IL1B-induced TNF mRNA decayed with a t½ of 30-40 min, and this was not affected by LMNA siRNA. However, silencing of ZFP36 significantly attenuated the loss of TNF mRNA (Fig. 4.9B).

This supports a post-transcriptional role for ZFP36 in the regulation of IL1B-induced TNF mRNA in A549 cells.

The effect of ZFP36 knock-down was also tested on IL1B plus dexamethasone-induced TNF mRNA and unspliced nuclear TNF RNA expression (Fig. 4.9A). Dexamethasone produced a significant attenuation of IL1B-induced TNF mRNA expression at all times. However, IL1B plus dexamethasone-induced TNF mRNA expression was not changed by ZPF36 knock-down.

Because the expression of IL1B-induced TNF mRNA was enhanced following ZFP36 silencing at

2 h, the actual percentage repression by dexamethasone was also increased at 2 h (Fig. 4.9A, right panels). Conversely, ZFP36 silencing did not produce any significant effect on unspliced nuclear

TNF RNA in the presence IL1B plus dexamethasone (Fig. 4.9A, lower panels).

105 A time (h) 1 2 6 ZFP36 siRNA 2 + + + + + + ZFP36 siRNA 1 + + + + + + LMNA siRNA + + + + + + Dex + + + + + + + + + time (h) 1 2 6 IL1B - + + + + + + - + + + + + + - + + + + + + NS ZFP36 siRNA 2 + + + ZFP36 ZFP36 siRNA 1 + + + LMNA siRNA + + + GAPDH IL1B+Dex + + + + + + + + + 300 *** *** *** ***  100 200 *** ** *** ***

*** * * *** 50

LMNA

siRNA)

siRNA) mRNA

100 Dexon

(%IL1B+

Relevant Effectof

(% IL1B (% TNF/GAPDH TNF/GAPDH 0 0 150  100 100 /U6

RNA 50

LMNA

siRNA) siRNA)

50 Dexon

(%IL1B+

Relevant Effectof

unTNF/U6 unTNF 0 IL1B (% 0

time post Act D (min) 45 B ZFP36 siRNA 2 + LMNA siRNA ZFP36 siRNA1 ZFP36 siRNA2 150 150 150 ZFP36 siRNA 1 + IL1B LMNA siRNA + 100 100 100 IL1B+siRNA IL1B + + + * 75 ** 50 50 * ** 50 *** **

mRNA 50 (% t =t (% 0) 0 TNF/GAPDH 0 0 25

0 20 40 60 0 20 40 60 0 20 40 60 mRNA (% t =t (%0) time (min) time (min) time (min) 0 TNF/GAPDH Figure 4.9 Effect of ZFP36 siRNA on IL1B-induced TNF mRNA expression. A, A549 cells were incubated with LMNA (control) or ZFP36-specific siRNAs for 24 h before treatment with IL1B (1 ng/ml) or IL1B plus dexamethasone (Dex, 1 μM) as indicated. Cells were harvested at 1, 2 or 6 h and total protein was prepared for western blot analysis of ZFP36 and GAPDH. Blots representative of at least 4 such experiments are shown. NS = nonspecific band. Cells were also harvested for real-time PCR analysis of TNF and GAPDH (middle panel) or unspliced nuclear (un) TNF RNA and U6 (lower panel). Data (N = 4) normalized to GAPDH or U6, were expressed as a percentage of LMNA siRNA plus IL1B-stimulated for each time and are plotted as means ± S.E. Significance was tested using ANOVA with a Newman-Keul multiple comparison test. Significance for specific comparisons are indicated. *, p < 0.05; **, p< 0.01; ***, p< 0.001. The effect of IL1B plus Dex expressed as a percentage of IL1B for each of the three individual siRNAs is plotted as a mean ± S.E. (right panels). B, A549 cells were incubated with LMNA or ZFP36- specific siRNAs for 24 h before being treated with IL1B (1 ng/ml) for 90 min. Actinomycin D (Act D, 10 μg/ml) was then added (t = 0) and the cells were harvested at the indicated times. RNA was extracted for real-time PCR analysis of TNF and GAPDH. Data (N = 4), normalized to GAPDH, are plotted as a percentage of t = 0 for each treatment as means ± S.E. Significance, relative to time matched IL1B-treated samples was tested using a paired t-test. *, p < 0.05; **, p < 0.01; ***, p < 0.001, (left panel). The effect of IL1B plus LMNA siRNA and IL1B plus ZFP36 siRNAs at 45 min post Act D treatment is plotted as a percentage of t = 0 for 45 min. Significance, relative to IL1B-treated samples was tested using ANOVA with a Dunnett's post test. **, p < 0.01, (right panel).

106 4.3.9 Effect of DUSP1 over-expression on ZFP36 and TNF mRNA expression

Since dexamethasone-induced DUSP1 represses IL1B-induced activation of MAPKs, and, as shown in this chapter, IL1B-induced ZFP36 is MAPK-dependent (Fig. 4.4B), the effect of DUSP1 over-expression on ZFP36 protein expression was also tested (382). The over-expression of

DUSP1 was confirmed by western blotting (Fig. 4.10). Control Ad-GFP virus was without any effect on DUSP1 expression. Inhibition of p38, JNK and ERK MAPK phosphorylation, and therefore activity, following DUSP1 over-expression are shown in the previous chapter and therefore were not reassessed in this chapter. ZFP36 expression was undetected in untreated cells and was not affected by DUSP1 or control GFP adenovirus (Fig. 4.10). IL1B-induced ZPF36 protein expression was unaffected by Ad-GFP, but Ad-DUSP1 almost completely inhibited ZFP36 expression at all times (Fig. 4.10).

In the case of TNF mRNA, both Ad-DUSP1 and control Ad-GFP had no effect on basal TNF mRNA expression (Fig. 4.10, lower panels). IL1B-induced TNF expression was not affected by

Ad-GFP. However, IL1B-induced TNF mRNA expression was significantly repressed by Ad-

DUSP1 at 1 h (Fig. 4.10, lower panels). This repression of TNF mRNA by DUSP1 over-expression was reduced at 2 h, and, surprisingly, by 6 h, there was no repression by Ad-DUSP1. These data suggest that even though the early induction of IL1B-induced TNF mRNA is MAPK-dependent, at longer time points, enhanced expression of DUSP1 by switching off MAPKs may overcome this effect on TNF mRNA expression.

107 time (h) 1 2 6 Ad-DUSP1 + + + + + + Ad-GFP + + + + + + IL1B - - - + + + - - - + + + - - - + + + DUSP1

GAPDH NS ZFP36 GAPDH 1.5 * 1.0 *

ZFP36 0.5

protein

GAPDH / 0.0 150 *** *** **

100 *** TNF

mRNA 50

GAPDH

/ (% of(% IL1B at each each at time) 0 Figure 4.10 Effect of DUSP1 over-expression on ZFP36 and TNF mRNA expression. A549 cells were either not infected or infected with Ad5-DUSP1 or Ad5-GFP (control) at a MOI of 10 for 24 h before IL1B treatment (1 ng/ml). Cells were harvested after 1, 2 or 6 h and total protein was prepared for western blot analysis of DUSP1, ZFP36 and GAPDH. Blots representative of at least 4 such experiments are shown. NS = nonspecific band. Following densitometric analysis for ZFP36, data (N = 4) were normalized to GAPDH and plotted as means ± S.E. (middle panels). Significance, using ANOVA with a Dunnett's post test is indicated. *, p < 0.05. Cells were also harvested for real-time PCR analysis of TNF and GAPDH (lower panels). Data (N = 4), normalized to GAPDH, are plotted as means ± S.E. Significance relative to time-matched IL1B and Ad5-GFP- treated samples, was tested using ANOVA with a Bonferroni post test. **, p< 0.01; ***, p< 0.001.

4.3.10 Effect of MAPK inhibitors on TNF expression

A549 cells were treated with the maximally effective concentrations of SB203580 (10 µM), U0126

(10 µM) or JNK inhibitor 8 (JNK-IN-8) (10 µM). TNF mRNA was strongly induced by IL1B at all times, and was not affected by any of the MAPK inhibitors at either 1 or 2 h post-IL1B treatment

(Fig. 4.11A). Conversely, IL1B-induced TNF mRNA expression was significantly enhanced by

SB203580 at 6 h (Fig. 4.11A). Since MAPK pathways are inhibited by both dexamethasone and

DUSP1 over-expression, the effect of the three MAPK pathways inhibitors in combination was also assessed. Combined inhibition of MAPKs produced a substantial and significant 81.2 ± 2.4%

108 loss of TNF mRNA following 1 h of IL1B treatment (Fig. 4.11A, right panel). However, by 2 h post-IL1B treatment, TNF expression was 54.7 ± 6.2% of IL1B control, and, at 6 h, there was an enhancement to 183.2 ± 22.2 % relative to IL1B treatment alone. Thus, even though TNF mRNA expression was attenuated by the combined inhibition of MAPKs at early time points, at later times, this repressive effect of MAPK inhibitors was both reversed and overcome.

The role of transcriptional and post-transcriptional mechanisms in the regulation of TNF by

MAPKs was assessed by measuring the accumulation of unspliced nuclear TNF RNA and TNF mRNA stability respectively. Accumulation of unspliced nuclear TNF RNA was not affected by individual MAPK inhibitors at 1 or 2 h (Fig. 4.11B). However, at 4 or 6 h post-IL1B treatment, while JNK-IN-8 showed a very modest effect, SB203580 and U0126 produced quite substantial, but not significant, enhancements of unspliced nuclear TNF RNA. In respect of combined MAPK inhibitors, IL1B-induced unspliced nuclear TNF RNA was significantly reduced at 1 h. Even though, there was no effect at 2 h, the accumulation of unspliced nuclear TNF RNA was enhanced by combined MAPK inhibitors at 4 and 6 h (Fig. 4.11B). These data suggest that the transcription of TNF at early times is MAPK-dependent. However, activation of MAPKs results in repression of TNF mRNA at later times.

In terms of the mRNA stability, TNF mRNA was very stable following 30 min IL1B treatment

(Fig. 4.2C). This effect was re-confirmed in Fig. 4.11C, where there was no overall loss of TNF mRNA over 60 min following actinomycin D chase. However, SB203580 produced an overall

~50% reduction in TNF mRNA post-actinomycin D treatment over this period. Pilot actinomycin

D chase experiments showed that SB203580 had no effect on the loss of TNF mRNA at either 2 or 4 h post-IL1B treatment (data not shown).

109 A 1.0 *** NT SB203580 UO126 JNK-IN-8 SB+UO+J8 IL1B 200 * 200 200 200 * 0.5

100 100 100 100

mRNA mRNA

** **

*** time) each **

TNF/GAPDH TNF/GAPDH 0.0 (% ofat IL1B 0 0 0 0 0 2 4 6 1 2 6 1 2 6 1 2 6 1 2 6 time (h) time (h) time (h) time (h) time (h)

B 1.5 NT SB203580 UO126 JNK-IN-8 SB+UO+J8 *

1.0 *** IL1B 400 400 400 400 RNA RNA 0.5 200 200 200 200

*** * *

unTNF/U6

unTNF/U6 each time) each

0.0 (% ofat IL1B 0 0 0 0 0 2 4 6 1 2 4 6 1 2 4 6 1 2 4 6 1 2 4 6 time (h) time (h) time (h) time (h) time (h)

C IL1B treatment time (min) 30 90 120 180 240 300 SB203580 + + + + + IL1B IL1B + + + + + + + + + + 200 IL1B + * SB203580 150 50 50 50 50 * mRNA 100 *

(% t =t 0) (% * 100

25 25 25 25 mRNA TNF/GAPDH * * 0 =t (%0) 50

0 20 40 60 TNF/GAPDH 0 0 0 0 0 time (min) Figure 4.11 Effect of MAPK inhibitors on TNF expression. A, A549 cells were either not treated (NT), treated with IL1B (1 ng/ml) or pre-treated with either UO126, SB203580, JNK inhibitor 8 (JNK-IN-8) or a combination of UO126, SB203580 plus JNK-IN-8 each at 10 µM for 30 min prior to IL1B stimulation. Cells were harvested after 1, 2 or 6 h for real-time PCR analysis of TNF and GAPDH. Data (N = 4) were normalized to GAPDH and are plotted as means ± S.E. The effect of IL1B + MAPK inhibitors for TNF mRNA is plotted as a percentage of IL1B at the indicated times. B, RNA from the experiments in A were subjected to real-time PCR analysis of unspliced nuclear (un) TNF RNA and U6. Data (N = 4) were normalized to U6 and are plotted as means ± S.E. The effect of IL1B + MAPK inhibitors for unspliced nuclear (un) TNF mRNA is plotted as a percentage of IL1B at the indicated times. For both A & B, significance, relative to either non-treated or IL1B-treated samples was tested using ANOVA with a Bonferroni post test. *, p < 0.05; **, p < 0.01; ***, p < 0.001. C, A549 cells were treated with IL1B (1 ng/ml) or pre- treated for 30 min with SB203580 at 10 µM prior to IL1B stimulation for 30 min. Following this (t = 0), actinomycin D (Act D, 10 μg/ml) was added and cells were harvested at the indicated times. RNA was extracted for real-time PCR analysis of TNF and GAPDH. Data (N = 4) normalized to GAPDH, are plotted as a percentage of t = 0 for each treatment as means ± S.E. (left panel). Data following 45 min of Act D treatment are shown (right panel). A549 cells were also treated with IL1B or IL1B + SB203580 at 10 µM for 90, 120, 180 and 240 min, prior to the addition of Act D (t = 0) for 45 min and analyzed as above (Right panels). Significance, relative to time matched IL1B-treated samples was tested using a paired t-test. *, p < 0.05.

Therefore, the loss of TNF mRNA was assessed following IL1B and IL1B plus SB203580 treatments for 90, 120, 180 and 240 min. Similar to 30 min of IL1B treatment, SB203580 significantly attenuated IL1B-induced TNF mRNA loss following 90 min IL1B stimulation.

110 However, SB203580 had no effect on IL1B-induced TNF mRNA loss following 120, 180 or 240 min treatments (Fig. 4.11C, right panel). These data suggest that at shorter IL1B treatment times, inhibition of p38 can destabilize and/or reduce nuclear RNA processing of TNF mRNA, but at longer treatment times there was no obvious stabilization of TNF mRNA. This could be attributed to the fact that in the absence of prior p38-dependent mRNA stabilization (occurring at 30 - 90 min post IL1B), there is no possibility of any subsequent p38-dependent mRNA destabilisation.

These data therefore support the possibility of existence of MAPK-dependent feed-forward mechanisms of TNF mRNA inhibition.

4.3.11 Effect of DUSP1 knock-down on ZFP36 protein and TNF mRNA expression

In the previous chapter, siRNA-mediated loss of DUSP1 enhanced the expression of IL1B-induced phosphorylated MAPKs at 1 h. Since IL1B-induced expression of both ZFP36 and TNF was

MAPK-dependent, the effect of DUSP1 loss on TNF and ZFP36 expression was also assessed.

DUSP1 protein was robustly induced by IL1B at 1 h, and this was substantially inhibited by the two DUSP1-targeting siRNAs (Fig. 4.12A). DUSP1 knock-down had no effect on TNF mRNA at either 1 or 2 h post-IL1B treatment (Fig. 4.12B, upper panel). However, DUSP1-targeting siRNAs produced a significant loss of IL1B-induced TNF mRNA expression at 6 h (Fig. 4.12B, upper panel). In respect of unspliced nuclear TNF RNA, DUSP1 knock-down had no effect at 1 or 2 h, and, at 6 h, a trend towards reduced expression was observed (Fig. 4.12B, lower panel). In the case of IL1B-induced ZFP36 expression, DUSP1 silencing produced a marginal or no effect at 1 h post- treatment (Fig. 4.12A). However, at 2 h following IL1B treatment, the expression of the upper

ZFP36 band was markedly enhanced by the two DUSP1-targeting siRNAs (Fig. 4.12A).

111 A time (h) 1 2 6 DUSP1 siRNA 2 + + + + + + DUSP1 siRNA 1 + + + + + + LMNA siRNA + + + + + + Dex + + + + + + + + + IL1B - + + + + + + - + + + + + + - + + + + + + NS DUSP1 GAPDH NS ZFP36 GAPDH B time (h) 1 2 6 DUSP1 siRNA 2 + + + + + + time (h) 1 2 6 + + + + + + DUSP1 siRNA 1 DUSP1siRNA 2 + + + LMNA siRNA + + + + + + + + Dex + + + + + + + + + DUSP1siRNA 1 + + + + - + + + + + + - + + + + + + - + + + + + + LMNA siRNA IL1B IL1B+Dex + + + + + + + + + 200 *** *** ** *** ***  ** ** 100 ** *** 100 ** 50

LMNA *

mRNA siRNA)

siRNA)

Dexon

(%IL1B+

Effectof Relevant

(% IL1B (% TNF/GAPDH 0 TNF/GAPDH 0 150 ** **  100 100 /U6

RNA 50

LMNA

siRNA) siRNA)

50 Dexon

(%IL1B+

Effectof Relevant

unTNF/U6 unTNF (% IL1B (% 0 0 Figure 4.12 Effect of DUSP1 knock-down on ZFP36 and TNF expression. A, A549 cells were incubated with LMNA (control) or DUSP1-specific siRNAs for 24 h before treatment with IL1B (1 ng/ml) or IL1B plus dexamethasone (Dex, 1 μM) as indicated. Cells were harvested at 1, 2 or 6 h and total protein was prepared for western blot analysis of DUSP1, ZFP36 and GAPDH. Blots representative of at least 7 – 9 such experiments are shown. NS = nonspecific band. B, Cells were treated as in A and harvested after 1, 2 or 6 h for real-time PCR analysis of TNF and GAPDH (upper left panel) or unspliced nuclear (un) TNF and U6 (lower left panel). Data (N = 4 - 8), normalized to GAPDH (for TNF mRNA) or U6 (for unspliced nuclear (un) TNF RNA), were expressed as a percentage of LMNA siRNA plus IL1B-stimulated for each time and are plotted as means ± S.E. Significance was tested using ANOVA with a Newman-Keul multiple comparison test. Significance for specific comparisons is indicated. **, p< 0.01; ***, p< 0.001. The percent IL1B plus dexamethasone/IL1B for the LMNA siRNA was also plotted and compared with that for each of the DUSP1-specific siRNAs using ANOVA with a Dunnett's post test. **, p< 0.01 (right panel).

The effect of DUSP1 knock-down in the presence of IL1B plus dexamethasone was also examined.

Co-treatment with dexamethasone attenuated IL1B-induced ZFP36 expression (Fig. 4.12A).

Conversely, the silencing of DUSP1, at 2 h, enhanced the expression of ZFP36 in the presence of

IL1B plus dexamethasone (Fig. 4.12A). In the case of TNF mRNA, while IL1B-induced TNF mRNA was repressed significantly by dexamethasone at all times, DUSP1-targeting siRNAs

112 produced no effect (Fig. 4.12B, upper panel). Nevertheless, the substantial loss of IL1B-induced

TNF mRNA that occurred at 6 h following DUSP1-targeting siRNAs resulted in a reduced percentage repression by dexamethasone at 6 h (Fig. 4.12B, upper right panel). Furthermore, the knock-down of DUSP1 revealed no significant effects on unspliced nuclear TNF RNA, suggesting that TNF transcription rate was not affected by DUSP1 loss in the presence of IL1B plus dexamethasone (Fig. 4.12B, lower panel). Since ZFP36 reduces TNF mRNA expression and stability (Fig. 4.9), the data presented in this section suggest that enhanced expression of ZFP36, at 2 h, following DUSP1 knock-down may potentiate the loss of TNF mRNA, at 6 h, in the presence of IL1B. Thus, the effect of ZFP36 silencing on TNF mRNA expression and stability was assessed in the presence of DUSP1 siRNAs.

4.3.12 Role of ZFP36 in the loss of TNF mRNA following DUSP1 knock-down

The role of ZPF36 was evaluated by transfecting the cells with DUSP1 and ZFP36-targeting siRNAs either alone or in combination. In each case, there was a significant knock-down of both

DUSP1 and ZFP36 (Fig. 4.13A). As shown in Fig. 4.9A, TNF mRNA expression was only marginally affected by the loss of IL1B-induced ZFP36 at 6 h (Fig. 4.13A, lower panel). In marked contrast, knock-down of DUSP1 alone produced a significant loss of TNF mRNA at 6 h (Fig.

4.13A, lower panel). This observation is in line with the data shown in Fig. 4.12B. However, additional loss of ZFP36, in DUSP1 inhibited cells, significantly reduced the repressive effect of

DUSP1 siRNAs (Fig. 4.13A, lower panel). These data confirm a role for ZFP36 in the enhanced loss of TNF mRNA following DUSP1 knock-down.

Since ZFP36 is an mRNA destabilizing protein, the stability of TNF mRNA was also assessed in the presence of DUSP1-targeting siRNAs. As shown in Fig. 4.2, IL1B-induced TNF mRNA was

113 decayed to ~50% following 45 min of actinomycin D treatment (Fig. 4.13B). DUSP1-targeting siRNAs significantly increased the loss of TNF mRNA (Fig. 4.13B) suggesting that in addition to enhancing ZFP36 expression at 2 h (Fig. 4.12A), DUSP1-targeting siRNAs also attenuate TNF mRNA stability. Thus, the effect of ZFP36 knock-down on TNF mRNA stability was tested in the presence and absence of DUSP1 siRNAs. The loss of TNF mRNA stability was significantly attenuated by ZFP36 siRNAs (Fig. 4.13B) and was consistent with previously shown effect of

ZFP36 knock-down on TNF mRNA stability in Fig. 4.9B. Thus, these data confirm that the stability of TNF mRNA is negatively regulated by ZFP36. In addition, the repressive effect of

DUSP1 siRNAs on TNF mRNA stability was significantly inhibited by the additional loss of

ZFP36 (Fig. 4.13B). In fact, the decay of TNF mRNA was similar to ZFP36 knock-down (Fig.

4.13B). These data therefore confirm that the enhanced decay of TNF mRNA following DUSP1 knock-down was ZFP36-dependent.

The effect of simultaneous knocking-down of DUSP1 and ZFP36 was assessed in the context of

IL1B and IL1B plus dexamethasone. In each case, the expression of both DUSP1 and ZFP36 was robustly inhibited by DUSP1- and ZFP36-targeting siRNAs (Fig. 4.13C). In respect of IL1B- induced TNF mRNA, combined knock-down of DUSP1 and ZFP36 produced a modest, but significant, repression at 1 h, no effect at 2 h, and a significant repression at 6 h (Fig. 4.13D, left panel). The effect of combined knock-down at 6 h was consistent with the data shown in Fig.

4.13A. IL1B-induced TNF mRNA was significantly repressed by dexamethasone at all times (Fig.

4.13D, left panel) and is similar to the data shown in Fig. 4.2A. There was no effect of combined

DUSP1 plus ZFP36 siRNAs on IL1B plus dexamethasone-induced TNF mRNA at 1 and 2 h (Fig.

4.13D, left panel). However, at 6 h, as a percentage of IL1B+dexamethasone+LMNA siRNA

114 A time (h) 1 C time (h) 1 ZFP36 siRNA 2 + + DUSP1 + ZFP36 siRNA 2 + + DUSP1 siRNA 2 + + ZFP36 siRNA 1 + + DUSP1 + ZPF36 siRNA 1 + + DUSP1 siRNA 1 + + LMNA siRNA + + Dex + + + IL1B - + + + + + + + IL1B - + + + + + + DUSP1 NS DUSP1 NS GAPDH GAPDH NS NS ZFP36 ZFP36 GAPDH GAPDH

150 6 h D * DUSP1 + ZFP36 siRNA 2 + + 100 * DUSP1 + ZPF36 siRNA 1 + + DUSP1 + ZFP36 siRNA 2 + ### ## + + ### ### LMNA siRNA DUSP1 + ZFP36 siRNA 1 +

mRNA 50

(%IL1B siRNA) +LMNA Dex + + + LMNA siRNA +

TNF/GAPDH 0 IL1B - + + + + + + IL1B+Dex + + + 150 *** 1 h * *** 100 B ** *** time post Act D (min) 45 100 * 50 ZFP36 siRNA 2 + + 50 ZFP36 siRNA 1 + + DUSP1 siRNA 2 + + 0 0 DUSP1 siRNA 1 + + 150 *** 2 h IL1B + + + + + + + *** 100 150 *** 100 *** ***

Relevant Relevant siRNA) 50

### mRNA

100 ### ### ### 50

 TNF/GAPDH

TNF/GAPDH 0 0 mRNA

50 EffectDexonof

# #

(% t =t 0) (% (%IL1B+LMNA (%IL1B+LMNA siRNA) TNF/GAPDH 150 *** 0 *** 6 h 100 100 ** *** IL1B (% *** 50 50 *** ** 0 0 Figure 4.13 Role of ZFP36 in the enhanced loss of TNF mRNA following DUSP1 knock- down. A, A549 cells were incubated with LMNA (control) or DUSP1- and/or ZFP36-specific siRNAs for 24 h before being treated with IL1B (1 ng/ml). While not indicated, LMNA siRNA was added to all the treatment groups except for the combination siRNA treatments to maintain a constant concentration of siRNAs across all the treatments. Cells were harvested after 1 h and total proteins were prepared for western blot analysis of DUSP1, ZFP36 and GAPDH. Blots representative of at least 4 such experiments are shown (top panel). NS = nonspecific band. Cells were also harvested after 6 h for real-time PCR analysis of TNF and GAPDH. Data (N = 4) normalized to GAPDH were expressed as a percentage of LMNA siRNA plus IL1B and plotted as means ± S.E. Significance was tested using ANOVA with a Newman-Keul multiple comparison test (lower panel). B, A549 cells were treated as in A before being treated with IL1B (1 ng/ml) for 2 h. Following this (t = 0), actinomycin D (Act D, 10 μg/ml) was added and cells were harvested after 45 min. RNA was extracted for real-time PCR analysis of TNF and GAPDH. Data (n = 6) normalized to GAPDH, are plotted as a percentage of t = 0 for each treatment as means ± S.E. For both A & B, significance between: LMNA control siRNA plus IL1B and each of the DUSP1 and/or ZFP36 targeting siRNAs plus IL1B is shown by #. Other comparisons are specifically indicated. */#, p < 0.05; ##, p< 0.01; ***/###, p< 0.001. C, A549 cells were incubated with LMNA (control) or DUSP1 and ZFP36-specific siRNAs for 24 h before being treated with IL1B (1 ng/ml) or IL1B plus dexamethasone (Dex, 1 μM) as indicated. Cells were harvested after 1 h and total proteins were prepared for western blot analysis of DUSP1, ZFP36 and GAPDH. Blots representative of at least 4 such experiments are shown. NS = nonspecific band. D, A549 cells were treated as in C and harvested after 1, 2 or 6 h for real-time PCR analysis of TNF and GAPDH. Data (N = 4) normalized to GAPDH were expressed as a percentage of LMNA siRNA plus (cont.…)

115 Figure 4.13 Continued. IL1B-stimulated for each time and are plotted as means ± S.E. Significance was tested using ANOVA with a Newman-Keul multiple comparison test. Significance for specific comparisons are indicated.*, p < 0.05; **, p< 0.01; ***, p< 0.001 (left panel).The percent IL1B plus dexamethasone/IL1B for the LMNA siRNA was also plotted and compared with that for each of the DUSP1 and ZFP36-specific siRNAs using ANOVA with a Dunnett's post test.*, p < 0.05 (right panel).

(16.7 ± 1.0 %), the expression of IL1B plus dexamethasone-induced TNF mRNA was significantly reduced (9.7 ± 0.6 & 8.3 ± 1.2 %) in the presence of the DUSP1 plus ZFP36 combined siRNAs

(Fig. 4.13D, left panel). Conversely, because each siRNA produced a significant loss of IL1B- induced TNF expression at 1 h, there was a reduction in the percentage repression by dexamethasone at 1 h (Fig. 4.13D, right panels). However, the percentage repression by dexamethasone at either 2 or 6 h was not changed (Fig. 4.13D, right panels).

4.3.13 Effect of MAPK inhibitors and DUSP1 over-expression in the regulation of TNF protein expression

IL1B-induced release of TNF into the supernatants was almost totally prevented by SB203580 and was significantly inhibited by both U0126 (74.0 ± 8.8%) and JNK-IN-8 (69.0 ± 20.2%) (Fig.

4.14A). Similarly, the combined inhibition of MAPKs also completely inhibited TNF release.

While the control, Ad-GFP virus, had no effect, DUSP1 over-expression significantly attenuated

TNF release into the supernatants (Fig. 4.14B). Thus, these data suggest that IL1B-induced TNF protein release is MAPK-dependent.

116 A time (h) 6 B JNK-IN-8 + + time (h) 6 UO126 + + Ad-DUSP1 + + SB203580 + + Ad-GFP + + IL1B - + + + + + IL-1B - - - + + + 200 *** 100 ** *** * ***

100 *** 50

TNF

TNF

(pg/ml)

(pg/ml)

release release 0 0 Figure 4.14 Effect of MAPK inhibitors and DUSP1 over-expression in the regulation of TNF protein expression. A, A549 cells were either not treated, treated with IL1B (1 ng/ml) or pre- treated with either UO126, SB203580, JNK inhibitor 8 (JNK-IN-8) or a combination of UO126, SB203580 plus JNK-IN-8 each at 10 µM for 30 min prior to IL1B stimulation. B, A549 cells were either not infected or infected with Ad5-DUSP1 or Ad5-GFP (control) at a MOI of 10 for 24 h before IL1B treatment (1 ng/ml). For both A & B, cells supernatants were harvested after 6 h for TNF release measurement. Data (N = 4) expressed in pg/ml are plotted as means ± S.E. Significance, using ANOVA with a Dunnett's post test (for A) or with a Bonferroni's multiple comparison test (for B) is indicated. *, p < 0.05; **, p< 0.01; ***, p< 0.001.

4.3.14 Effect of DUSP1- and ZFP36-targeting siRNAs in the regulation of TNF protein expression

To assess the role of ZFP36 and DUSP1 in regulating TNF protein expression, TNF release was measured from the experiments shown in Figs. 4.9 and 4.12 respectively. Whereas a control siRNA targeted to LMNA had no effect on IL1B-induced TNF release at 6 (Fig. 4.15), knock-down of

ZFP36 significantly enhanced the release of TNF following IL1B treatment (Fig. 4.16A) and is consistent with the enhanced expression of TNF mRNA observed at 2 h in Fig. 4.9.

time (h) 6 Figure 4.15 Effect of LMNA siRNA on TNF protein LMNA siRNA + + expression. A549 cells were incubated with or without Dex + + LMNA-specific siRNA. After 24 h, cells were treated with IL1B - + + + + IL1B (1 ng/ml) or IL1B plus dexamethasone (1 μM) (Dex) 150 *** *** ) as indicated. Cells supernatants were harvested after 6 h for

100 TNF release measurement. Data (N = 4) expressed in pg/ml TNF

pg/ml 50 (%IL1B

release release are plotted as means ± S.E. Significance, using ANOVA 0 with a Bonferroni's multiple comparison test is indicated. *, p < 0.05; **, p< 0.01; ***, p< 0.001.

117

A B time (h) 6 time (h) 6 ZFP36 siRNA 2 + + time (h) 6 DUSP1 siRNA 2 + + time (h) 6 ZFP36 siRNA 1 + + ZFP36 siRNA 2 + DUSP1 siRNA 1 + + DUSP1 siRNA 2 + LMNA siRNA + + LMNA siRNA + + ZFP36 siRNA 1 + Dex + + + DUSP1 siRNA 1 + Dex + + + LMNA siRNA + LMNA siRNA + IL1B - + + + + + + IL1B+Dex + + + IL1B - + + + + + + IL1B+Dex + + + 60 * 12 **

**  ***  100 40 *** *** 8 ** 100

*** TNF

50 TNF 4 50

(pg/ml) (pg/ml)

20 release

release release

Dexon siRNA)

Dexon siRNA)

release

release release

Relevant Effectof

Effectof Relevant (% IL1B (% 0 IL1B (% 0 0 0 C time (h) 6 time (h) 6 DUSP1 siRNA 2 + + + + DUSP1 siRNA 1 + + DUSP1 + ZFP36 siRNA 2 time (h) 6 ZFP36 siRNA 2 + + DUSP1 + ZPF36 siRNA 1 + + DUSP1 + ZFP36 siRNA 2 + ZFP36 siRNA 1 + + LMNA siRNA + + DUSP1 + ZFP36 siRNA 1 + + + + IL1B - + + + + + + + Dex LMNA siRNA + IL1B - + + + + + + 40 IL1B+Dex + + + * *** *** 80 **  *** *** 100 20 ** ***

TNF 40 50

TNF

siRNA)

release (pg/ml)

Dexon

release (pg/ml)

release release

Relevant Effectof 0 0 IL1B (% 0

Figure 4.16 Effect of DUSP1- and ZFP36- targeting siRNAs in the regulation of TNF protein expression. A, B & C, cells were incubated with LMNA (control) or DUSP1-and/or ZFP36- specific siRNAs for 24 h before being treated with IL1B (1 ng/ml) or IL1B plus dexamethasone (Dex, 1 μM) as indicated. Cells supernatants were harvested after 6 h for TNF release measurement. Data (N = 4) expressed in pg/ml are plotted as means ± S.E. Significance, using ANOVA with a Newman-Keul multiple comparison test is indicated. *, p < 0.05; **, p< 0.01; ***, p< 0.001 (left panels). In each case, the percent IL1B plus dexamethasone/IL1B for the LMNA siRNA was also plotted and compared with that for each of the DUSP1- and ZFP36-specific siRNAs using ANOVA with a Dunnett's post test (right panels). For C (left panel): LMNA siRNA was added in all the treatment groups except for the combination siRNA treatments in order to maintain a constant siRNA concentration across all the treatments.

However, DUSP1 siRNAs were without any obvious effect on TNF release (Fig. 4.16B). The effects of combined knock-down of ZFP36 and DUSP1 were also examined (Fig. 4.16C). Thus, as shown in Fig. 4.16A, IL1B-induced TNF release was augmented by ZFP36 siRNA, whereas

DUSP1 siRNA was without any marked effect (Fig. 4.16C). In the presence of combined ZFP36 plus DUSP1 siRNA, TNF release was higher than following DUSP1 knock-down alone, but in each case this was not significantly different from the effect of ZFP36-targeting siRNAs alone

(Fig. 4.16C, left panel). In the presence of IL1B plus dexamethasone, TNF release was not

118

significantly changed by the knock-down of ZFP36 and DUSP1 either individually (Fig. 4.16A &

B) or in combination (Fig. 4.16C, middle panel). Furthermore, the percentage repression by dexamethasone was also unaltered by either treatment (Fig. 4.16A, B & C, right panels).

4.4 Discussion

In the current chapter, the role of phosphatase, DUSP1, and the mRNA destabilizing protein,

ZFP36, in the regulation of TNF mRNA, as a model ARE-containing transcript, was examined. In this context, IL1B-induced ZFP36 reduces TNF mRNA stability and exerts feed-forward inhibition on TNF mRNA and protein expression in A549 cells. Thus, the data presented here clearly explain why mechanistically simple explanations, not taking into account overall network function, are unlikely to provide a sufficient explanation for repression by glucocorticoids.

In A549 cells, IL1B-induced ZFP36 expression was MAPK-dependent, and primarily mediated by the p38 MAPK. In contrast, inhibition of MAPKs, in particular p38 MAPK, enhanced IL1B- induced TNF mRNA expression and was appeared to be primarily associated with enhanced transcription. These data therefore suggest the existence of, as yet uncharacterised, MAPK- dependent mechanisms that negatively regulate TNF transcription. In terms of the stability of TNF mRNA, following 30 min of IL1B stimulation, TNF mRNA appeared to be stable within the 45 min of actinomycin D chase. This contrasts with longer IL1B treatment times, where ZFP36 is expressed and TNF mRNA decayed to 50% within 30 - 40 min of actinomycin D treatment. In the presence of the p38 MAPK inhibitor, SB203580, the apparent mRNA stabilisation, observed following 30 min of IL1B treatment, did not occur and at longer times there was no destabilisation of what was presumably an already unstable transcript. These data and interpretation are consistent

119

with previous findings in macrophages where TNF mRNA was stabilized transiently following 30

- 60 min of LPS treatment, but this stabilisation was lost within 2 h post-LPS treatment (383).

Similarly, a role for p38 MAPK in the stabilisation of other ARE-containing mRNAs, including

IL8, PTGS2 and TNF has also been indicated (227, 376, 384-386).

In the previous chapter, IL1B-induced MAPK phosphorylation was enhanced following DUSP1 knock-down at 1 h (Fig. 4.17A). Equally, as shown in this chapter, silencing of DUSP1, by enhancing the expression of ZFP36 protein at 2 h, subsequently increased the loss of the ARE- containing TNF mRNA at 6 h (Fig. 4.17A). This regulatory loop is therefore consistent with a previously indicated role for p38 MAPK in inducing the expression and activity of ZFP36 to destabilize ARE-containing mRNAs, such as TNF (309, 387). In addition, the observed attenuation of IL1B-induced TNF mRNA following DUSP1 silencing was not primarily associated with reduced TNF transcription, although some loss of transcription was evident at 6 h. Rather, the enhanced loss of TNF mRNA following DUSP1 knock-down was inhibited by further knock-down of ZFP36. This was mainly attributed to reduced TNF mRNA stability, an effect that was also blocked by ZFP36 silencing. Thus, these data directly confirm that loss of DUSP1 enhances ZFP36 expression to increase negative feed-forward control and attenuate TNF mRNA stability (Fig.

4.17A). Since the loss of ZFP36 did not completely prevent all the effects of DUSP1 knock-down on TNF mRNA expression, the possibility of additional non-ZFP36-dependent, but MAPK- dependent, regulation of TNF mRNA expression is suggested. In this regard, AREs are targeted by a number of RNA binding proteins, including HuR (ELAVL1), AUF1 (HNRNPD) and KHSRP

(220, 221, 223). These factors, by competing for ARE binding, inhibit the biosynthesis of

120

inflammatory cytokines and chemokines (214, 215). In this regard, mice deficient in AUF1 show exaggerated inflammatory responses following endotoxin activation due to the enhanced expression of TNF and IL1B (388). Similarly, the translation of TNF in macrophages was profoundly inhibited by the over-expression of HuR (389). Thus, TNF expression can also be negatively regulated by non ZFP36-dependent mechanisms.

The above findings highlight a number of key points in respect of the regulation of TNF mRNA expression by glucocorticoids. The IL1B-induced accumulation of unspliced nuclear TNF RNA was partially attenuated by dexamethasone suggesting that dexamethasone targets TNF transcription. This effect was largely reflected in TNF mRNA, which was modestly, but significantly, repressed by dexamethasone at all times. While transcriptional repression was clearly important, actinomycin D chase experiments showed that the loss of TNF mRNA was also increased by dexamethasone. Thus, as has been observed for other inflammatory genes, transcriptional and post-transcriptional processes may both contribute to the repressive effects of glucocorticoids on TNF mRNA (254, 390). Since MAPK pathways play an important role in transcriptional and post-transcriptional processes, and are strongly repressed by dexamethasone, the implication of MAPK pathways in these processes points to possible repressive mechanisms

(227, 281, 291, 293, 298, 300). However, with co-treatment, there is little repression of MAPK phosphorylation by dexamethasone at 30 min post-IL1B treatment (Fig. 4.17B) (114, 283, 382).

Therefore, this is unlikely to be responsible for the observed repression of TNF mRNA by dexamethasone at 30 min. In this context, since dexamethasone-induced repression of TNF mRNA was unaffected in the presence of the translational blocker, cycloheximide, effects due to

121

glucocorticoid-induced gene expression can be ruled out, and this therefore supports a role for direct transrepression by NR3C1 (273) (see chapter six, section 6.4 for further details).

In the case of ZFP36, since ZFP36 expression was MAPK-dependent, dexamethasone produced a modest inhibition of IL1B-induced ZFP36 expression at 1 and 2 h (Fig. 4.17B). This was consistent with previously shown repressive effect of dexamethasone on IL1B- and LPS-induced ZFP36 in

A549 cells and macrophages, respectively (308, 391). Contrary to this, a number of studies report that ZFP36 can also be induced by glucocorticoids (306, 308, 317, 318). This also holds true for the data shown in this chapter. However, the inducibility of ZFP36 by glucocorticoids in A549 cells was relatively modest. In any case, this glucocorticoid-mediated induction of ZFP36 may allow the expression of ZFP36 to be restored or maintained at longer times (i.e. 6 h and thereafter) following IL1B plus dexamethasone treatment (Fig. 4.17B). Nevertheless, loss of MAPK activity by dexamethasone-dependent effector mechanisms will reduce ZFP36-mediated negative feed- forward control (Fig. 4.17B). While this should enhance TNF mRNA expression at longer times, this does not happen in the presence of glucocorticoids. In fact, there was more than 50% repression of TNF mRNA by dexamethasone at 6 h post-IL1B treatment (Fig. 4.2A). In addition, silencing of DUSP1 or ZFP36, either alone or in combination, did not affect the repression of TNF mRNA by dexamethasone at any time. Therefore, a role for additional, non-DUSP1/non-ZFP36- dependent effector mechanisms of repression by glucocorticoids is predicted (254, 392) (Fig.

4.17B).

122

The above data document the existence of the regulatory mechanisms illustrated in Fig. 4.1 in respect of TNF mRNA expression and a similar scenario was also evident for protein release. In

A549 cells, TNF protein was released rapidly into the supernatants following peak mRNA expression. This reached ~100 pg/ml (~50 pg/106 cells), a level that is within the range (30 – 900 pg/106 cells) that are produced by mast cells, which are a physiologically relevant source of TNF

(393-397). Furthermore, in human airway epithelial cells, low nanomolar levels of TNF were sufficient to produce maximal responses, including cytokine release and activation of NF-κB

(398). In addition, low picomolar levels (~10 nM) of TNF were also capable of producing responses, particular in the context of co-stimulants such as histamine (398). Thus, the low amounts of TNF produced by A549 and primary HBE cells could be sufficient to elicit a biological response. In addition, since membrane-tethered TNF is biologically active, the expression of the membrane tethered uncleaved, ~25 kDa, TNF was also examined (399-401). Uncleaved TNF expression induced by IL1B was similar in both A549 and primary HBE cells, and in each case, dexamethasone produced a significant repression of IL1B-induced ~25 kDa TNF. Furthermore, in

A549 cells, while the expression of the IL1B-induced uncleaved TNF had largely diminished by

6 h, cell-associated TNF was maintained at a relatively higher level. This membrane tethered and/or cell-associated TNF may play key roles in the rapid activation of resident or infiltrating cells. However, the analysis of the functional relevance are beyond the scope of the current chapter.

Even though inhibition of MAPK significantly inhibited TNF protein expression, DUSP1 silencing had no effect on TNF release. In contrast, TNF release was significantly increased by ZFP36 silencing and indicates a key role for ZFP36 in the negative regulation of TNF translation. While

123

the exact mechanism of this repression is unclear, it may involve ZFP36-dependent loss of the polyA tail, which, by decreasing the translation efficiency and mRNA stability can reduce TNF biosynthesis (215, 377, 402). Thus, while transient increases in MAPK activation following

DUSP1 silencing promoted MAPK-dependent TNF synthesis, the corresponding enhancement of

ZFP36 expression following DUSP1 loss could counterbalance this effect resulting in no net effect of DUSP1 silencing on TNF protein release (Fig. 4.17A). Because there was enhanced TNF protein release following ZPF36 loss, the scheme depicted in Fig. 4.1 holds true for TNF protein synthesis.

In the case of repression by dexamethasone, IL1B-induced TNF protein was almost completely repressed by dexamethasone. Since TNF mRNA was partially repressed, this implies that translational and/or post-translational mechanisms may play an important role in the inhibition of

TNF expression by glucocorticoids, and such mechanisms have been previously suggested (403).

Furthermore, knock-down of ZFP36 or DUSP1 alone had no effect on the level of TNF released by IL1B plus dexamethasone. These data suggest that neither DUSP1 nor ZFP36 plays any overriding role in glucocorticoid-mediated repression of TNF protein in this model. Moreover, the observed enhancement in the percentage repression of TNF release by dexamethasone was mainly attributed to the fact that loss of ZFP36 enhanced the level of IL1B-induced TNF release (Fig.

4.16A). Similarly, combined knock-down of both ZFP36 and DUSP1 did not affect IL1B plus dexamethasone-induced TNF release (Fig. 4.16C middle panel). This suggests that additional glucocorticoid-induced repressive mechanisms must account for the repression of TNF release by dexamethasone (Fig. 4.17B).

124

The findings shown in this chapter, in respect of the effects of IL1B and dexamethasone on

DUSP1, ZFP36 and TNF expression, were similar in both A549 and primary HBE cells. Thus, the regulatory mechanisms that are proposed for A549 cells are likely to be relevant to primary human airways epithelial cells. In addition, recent data in other cell lines, including primary human airway smooth muscle cells, as well as in mouse models also support that ZFP36 plays an important role in the inhibition of ARE-containing inflammatory transcripts, such as TNF (Fig. 4.1) (404, 405).

In particular, the interaction between DUSP1-mediated feedback control and ZFP36-mediated feed-forward control shows how modulation of a single signalling component can lead to opposing effects on gene expression at different times. Finally, ZFP36 may not completely explain the effects of DUSP1 silencing. Thus, ARE-containing inflammatory gene expression could be negatively regulated by a number of additional non ZFP36-dependent (but MAPK-dependent) mechanisms. Such data add to the complexity of the system and confirm the need for detailed modelling to assess therapeutic strategies (406). In respect of the anti-inflammatory effects of glucocorticoids, these data suggest that neither DUSP1 nor ZFP36 play a dominant repressive role on TNF expression. Thus, while negative feed-forward control may be down-regulated by the early effects of glucocorticoids, at longer times, additional repressive mechanisms by glucocorticoids must also exist in order to maintain the repression of TNF.

125

A B IL1B treatment (No DUSP1) IL1B + glucocorticoid IL1B IL1B Other GC effectors MAPK GC DUSP1 MAPK Feed-forward Feed-forward GC control by ZFP36 control (ZFP36) TNF gene TNF gene Other GC expression expression effectors 150 DUSP1 siRNA 700 Phosho-p38 Phospho-p38 200 Phospho-p38 100 150 350 50 100 50 0 0 200 500 DUSP1 protein 0 DUSP1 protein ZFP36 protein

200 100 250

Gene/GAPDH % 1 h h IL1B+Lsi 1 % 100 0 0 ¼ ½ 1 2 6 150 150 0 ZFP36 protein ZFP36 protein 150

100 100 IL1B+Dex TNF mRNA h IL1B 1 % 100 Gene/GAPDH 50 50

50 0 ofat (% each time) IL1B 0 1 2 6

300 % IL1B+Lsi % TNF/GAPDH 0 TNF mRNA 100 0 2 4 6 200 TNF mRNA 50 time (h) 100 NT IL1B IL1B+siRNA 0 0 0 2 4 6 1 2 4 6 time (h) time (h) NT IL1B IL1B+Dex Figure 4.17 Regulation of TNF gene expression following DUSP1 silencing and glucocorticoid treatment. Schematics representing possible regulatory networks are shown along with summarized data from the current and a prior chapter (chapter three). A, With reduced DUSP1 expression there is a reduced negative feedback control of MAPKs leading to enhanced MAPK activity (at 1 h) (bold) at the times where DUSP1 expression would have been elevated. This could promote enhanced TNF expression. However, enhanced MAPK activity enhances expression of ZFP36 (at 2 h) (bold), which may decrease TNF expression (at 6 h). Actual expression of TNF is dependent on the temporal interplay of these competing regulatory processes. B, In the presence of glucocorticoid co-treatment, DUSP1 expression is enhanced (bold) and promotes inactivation of MAPKs (grey). Loss of MAPK activity reduces the expression of feed-forward (cont...)

126

Figure 4.17 Continued. negative control genes, such as ZFP36. However, ZFP36 is also modestly induced by glucocorticoid alone (in brackets) and this may help maintain feed-forward control. Given the net loss of IL1B-induced ZFP36 expression in the presence of glucocorticoids (at 1 and 2 h), additional glucocorticoid effectors must exist to maintain the repression of TNF (at later times). Repression of MAPKs is also maintained by non-DUSP1-dependent mechanisms. Positive signalling/expression (blue) is represented by arrows. Negative effects are indicated by (red) lines ending in a T-bar. Time course expression of phospho-p38, DUSP1 and ZFP36 protein and TNF mRNA in the presence of IL1B and IL1B plus DUSP1 siRNA (A) or IL1B plus dexamethasone (B) is shown. The protein expression data for DUSP1 and ZFP36 were generated following densitometric analysis of western blots showed in Fig. 4.3. For phospho-p38, western blots were taken from chapter three. For phospho-p38, DUSP1, ZFP36 and TNF, at each time, the effect of IL1B plus Dex expressed as a percentage of IL1B is plotted as a mean (B, right panels).

127

Chapter 5: Maintenance of IRF1 by DUSP1 enhances IRF1-dependent gene expression: Implications for glucocorticoid therapy

5.1 Rationale

In previous work by King et. al., the effect of dexamethasone on 39 IL1B-induced inflammatory genes was analysed in A549 cells (273). Based on the peak of their mRNA induction these genes were placed into 3 different groups. Genes with a peak in mRNA expression occurring at 1 or 2 h post-IL1B treatment were called ‘early-phase’ genes. In contrast, genes that had a peak of mRNA expression at 6 h and less than 50% of this peak at 1 or 2 h were defined as ‘late-phase’ genes.

Finally, genes that had a peak of mRNA expression at 2, 6 or 18 h post-IL1B treatment, and did not fall into neither of the above two categories, were referred to as ‘intermediate’ genes (273). In addition, in chapter three, while investigating the role of DUSP1 as a regulator of dexamethasone- induced repression of inflammatory genes, DUSP1 was over-expressed. As expected, this led to the repression of many inflammatory genes, including CSF2 (GM-CSF), IL8 and PTGS2. In contrast, in this chapter, many other IL1B-induced inflammatory genes, including IRF1, ICAM1,

CFB, CXCL10, IFIT1, MX1, UBD etc. were substantially enhanced by DUSP1 over-expression.

With the exception of IRF1 and ICAM1, the majority of these, IL1B-induced, DUSP1-enhanced genes were late-phase genes (273), and may therefore require the prior expression of other gene products, for example a transcription factor, for their expression. This concept was further confirmed by the effect of protein synthesis inhibitor, cycloheximide, where inhibition of de novo protein synthesis led to the inhibition of these genes by IL1B (273). Thus, the enhanced expression of IRF1, a transcription factor, by DUSP1 could be responsible for the increased expression of late-phase genes. Furthermore, the DUSP1-mediated increase in IRF1 expression could contribute

128

to a lack of glucocorticoid sensitivity for such genes. Indeed IRF1 has been previously implicated in playing a role in glucocorticoid insensitivity (210, 407).

5.2 Hypothesis

The hypothesis that up-regulation of DUSP1 by dexamethasone helps to maintain IL1B-induced

IRF1 expression and thereby sustain the expression of downstream IRF1-dependent genes is tested.

5.3 Results

5.3.1 Effect of DUSP1 over-expression on inflammatory gene expression

A549 cells were treated with IL1B, for 6 h, in the presence of a DUSP1-expressing adenovirus,

Ad-DUSP1, or a control virus. While control virus had little or no effect on IL1B-induced inflammatory gene expression, many inflammatory genes, including CCL4, CCL20, CSF2,

CXCL1, CXCL3, CXCL5, IL8, PTGS2 and others showed a significant repression by Ad-DUSP1

(Fig. 5.1A). This is consistent with the prior observations in chapter three. Conversely, IL1B- induced BIRC3, CCL2, CCL5, IFNGR1, IL6, LAMB3, MAP3K8, NFKB2, NFKBIZ and

TNFAIP3 mRNAs were partially inhibited by DUSP1 over-expression (Fig. 5.1A). However, a number of mRNAs, including APOL6, CFB, CMPK2, CXCL10, HELZ2, ICAM1, IFIT1, IFIT3 isoforms 1 & 2, IRF1, MX1, STAT5A, TLR2 and UBD revealed significantly enhanced IL1B- induced expression following DUSP1 over-expression (Fig. 5.1A). Additionally, EFNA1, GOS2 and TNF mRNAs were also enhanced by 1.34 ± 1.4, 1.47 ± 0.8 and 2.22 ± 1.7 fold respectively by DUSP1 over-expression, but this effect did not attain significance (Fig. 5.1A).

129

Gene p Gene A NS IL1B IL1B+Ad5-DUSP1IL1B+Ad5-Null B D F NS IL1B+LMNAsiIL1B+IRF1si1IL1B+IRF1si2 p PI3 *** time (h) 18 ICAM1 CCL5 *** *** TFF1 *** STAT5A CXCL5 *** *** CSF2 *** Ad-GFP + + CCL20 CXCL3 *** Ad-DUSP1 + + PTGS2 *** 1.5 CSF2 CXCL5 *** IL1B - - - + + + TFF1 IL8 *** 1.0 TNF 2000 *** *** CXCL1 *** EFNA1 CCL4 ** mRNA 0.5 TLR2 BCL2A1 * 1000 FAM129A *** 0.0 ICAM1

Gene /GAPDH

CCL20 ** (pg/ml) BIRC3 Release Release

CXCL10 1 2 6 18 IL1B *** 0 TNFAIP3 OLR1 *** 8 GOS2 CSF3 *** ISG20 ** C MAP3K8 SOD2 *** 1000 4 PI3 NFKBIA *** IRF1 CCL2 IL32 * mRNA IL6 MAP3K8 500 BIRC3 0 STAT5A IFNGR1 Gene /GAPDH PTGS2 mRNA 1 2 6 18 CCL2 0 CFB * * CCL5

IRF1/GAPDH time (h) UBD * * LAMB3 1 2 6 18 IFIT3iso1 * * IL6 time (h) TNFAIP3 APOL6 IFIT3iso2 HELZ2 * ** NFKB2 TLR2 IFIT3iso1 APOL6 *** *** NFKBIZ NT IL1B CFB UBD IFIT1 *** *** EFNA1 CMPK2 HELZ2 MX1 *** *** ICAM1 ** IFIT1 MX1 CXCL10 * ** GOS2 CXCL10 IFIT3iso1 ** time (h) 1 1 2 6 18 18 CMPK2 *** *** HELZ2 * IFIT3iso2 *** *** APOL6 *** IL1B - + + + + - CFB *** E TNF IRF1 time (h) 2 MX1 ** IRF1 siRNA 2 + 0 100 400 STAT5A ** GAPDH UBD *** IRF1 siRNA 1 + CMPK2 *** LMNA siRNA + % IL1B IRF1 *** 4 *** IL1B - + + + + IFIT1 *** TLR2 *** 2 IRF1 IFIT3iso2 *** CXCL10 IRF1 *** protein /GAPDH 0 GAPDH

0 100 400 % IL1B Figure 5.1 Effect of DUSP1 over-expression, IL1B and IRF1-targeting siRNA on inflammatory gene expression. A, A549 cells were either not infected or infected with Ad5- DUSP1 or Ad5-Null/Ad5-GFP (control) at a MOI of 10 for 24 h before IL1B treatment (1 ng/ml). Cells were harvested after 6 h for real-time PCR analysis of the indicated genes and GAPDH. Data (N = 4), normalized to GAPDH, were expressed as a percentage of IL1B-stimulated cells and plotted as means ± S.E. Significance relative to IL1B-treated samples, was tested by ANOVA with a Dunnett's post-test. *, p < 0.05; **, p< 0.01; ***, p< 0.001. B, Cells were treated as in A and the supernatants harvested after 18 h for CXCL10 release measurement. Data (N = 4) expressed in pg/ml are plotted as means ± S.E. Significance, relative to time-matched IL-1B and Ad5-GFP- treated samples, was tested by ANOVA with a Bonferroni post test. *, p < 0.05; **, p< 0.01; ***, p< 0.001. C, A549 cells were either not treated (NT) () or stimulated with IL1B (1 ng/ml) () as indicated. Cells were harvested after 1, 2, 6 or 18 h for real-time PCR analysis of IRF1 and GAPDH. Data (N = 4) were normalized to GAPDH and plotted as means ± S.E. (upper panel). Cells were also harvested at the times indicated for western blot analysis of IRF1 and GAPDH (lower panel). Representative blots are shown. Following densitometric analysis, data (N = 4) were normalized to GAPDH and plotted as means ± S.E. Significance, using ANOVA with a Dunnett's post test is indicated. ***, p< 0.001. D, Cells were treated as in C and harvested after 1, 2, 6 or 18 h for real-time PCR analysis of indicated genes and GAPDH. Data (N = 4) were normalized to GAPDH and plotted as means. E, A549 cells were incubated with LMNA (control) or (cont.…)

130

Figure 5.1 Continued. IRF1-targeting siRNAs. After 24 h, cells were treated with IL1B (1 ng/ml) as indicated. Cells were harvested after 2 h for western blot analysis of IRF1 and GAPDH. Blots representative of at least 4 such experiments are shown. F, A549 cells were treated as in E and harvested after 6 h for real-time PCR analysis of indicated genes and GAPDH. Data (N = 4), normalized to GAPDH, were expressed as a percentage of IL1B-stimulated cells and plotted as means ± S.E. For each IRF1-targeting siRNA, significance relative to IL1B+LMNA siRNA- treated samples, was tested by ANOVA with a Dunnett's post test. *, p < 0.05; **, p< 0.01; ***, p< 0.001.

While adenoviral over-expression of DUSP1 significantly enhanced IL1B-induced CXCL10 release, the control, Ad-GFP virus, was without any effect (Fig. 5.1B).

5.3.2 Kinetics of IL1B-induced inflammatory mRNAs

Following IL1B treatment, IRF1 mRNA was highly induced at 1 h, reached a peak 2 h post- stimulation, and then declined steeply towards basal levels over the following 4 h (Fig. 5.1C, upper panel). This expression kinetic correlated closely with IRF1 protein expression, which also peaked at 2 h before declining rapidly by 6 h following IL1B treatment (Fig. 5.1C, lower panels). The mRNA expression of inflammatory genes, showing significant enhancement by DUSP1 over- expression, was also tested. The expression of ICAM1 and STAT5A was modestly induced at 1 h with a peak in expression at 2 - 6 h, prior to declining at later times (Fig. 5.1D, upper panel). These data contrast with the effects on APOL6 and other inflammatory genes, which were all modestly induced by IL1B at 1 and 2 h, but showed a peak in expression at 6 h (Fig. 5.1D, lower panel).

CFB mRNA was not induced until 6 h and revealed a further increase in expression at 18 h post-

IL1B treatment (Fig. 5.1D, lower panel). Thus, with the exception of IRF1, STAT5A and ICAM-

1, the majority of these IL1B-induced, DUSP1-enhanced genes fell into the "late" group of genes that were induced by IL1B. This suggests that their mRNA expression requires, or depends on, the prior expression of another gene product, most likely a transcription factor. Since IRF1 has an

131

“early” kinetic of expression, and is a transcription factor, a role in the induction of downstream genes was tested.

5.3.3 Identification of IRF1-dependent mRNAs induced by IL1B

A549 cells were treated with two independent IRF1-targeting or LMNA (control)-targeting siRNAs prior to stimulation with IL1B. IL1B-induced IRF1 protein expression was robustly inhibited by both of the IRF1-targeting siRNAs (Fig. 5.1E). The control LMNA siRNA, was without effect, (Fig. 5.1E). After harvesting at 6 h, qPCR was carried out for the IL1B-induced genes that were significantly increased by DUSP1 over-expression, plus a number of other genes showing no effect, or repression, following DUSP1 over-expression. All the late-phase mRNAs,

(Fig. 5.1D, lower panel) except TLR2, that had previously shown significant enhancement by

DUSP1-over expression were significantly reduced by IRF1 silencing (Fig. 5.1F). In contrast, the early response genes, STAT5A and ICAM1, were not down-regulated by IRF1 silencing and therefore were not IRF1-dependent (Fig. 5.1F). Additionally, the expression of those genes that were either not affected, or repressed, by DUSP1 over-expression, such as CSF2, TFF1, EFNA1,

BIRC3, TNFAIP3 etc., was also not changed by IRF1 knock-down (Fig. 5.1F). However, the

IL1B-induced expression of CCL5 and CXCL5 showed significant 52.08 ± 5.6 and 20.65 ± 8.2 fold enhancements respectively compared to IL1B plus LMNA siRNA treated cells following

IRF1 silencing (Fig. 5.1F). This suggests that the expression of these mRNAs may be negatively regulated by IRF1, or an IRF1-dependent process. Given a role for IRF1 in the expression of IL1B- induced late-phases genes that were enhanced by DUSP1 over-expression, the effects of DUSP1 over-expression and/or MAPKs inhibition on IRF1 expression were also examined.

132

5.3.4 Effect of DUSP1 over-expression on IL1B-induced IRF1 expression

A549 cells were treated with Ad-DUSP1 or control Ad-GFP adenovirus in the presence and absence of IL1B. The over-expression of DUSP1 was confirmed by western blotting (Fig. 5.2A).

IRF1 protein was only modestly detectable at 1 h post-IL1B treatment. At 2 and 4 h, IRF1 was highly induced by IL1B, but there was no effect of DUSP1 over-expression. Conversely, while

IL1B-induced IRF1 expression had declined by 6 h, over-expression of DUSP1 appeared to maintain IL1B-induced IRF1 expression. Control Ad-GFP virus had no effects on IL1B-induced

IRF1 expression (Fig. 5.2A).

As described in Fig. 5.1C, IRF1 mRNA was induced by IL1B at 1 h, reached a peak at 2 h, and declined sharply by 6 h (Fig. 5.2B). While IL1B-induced IRF1 mRNA was unaffected by the control GFP virus, Ad-DUSP1 significantly (45.9 ± 8.8%) attenuated IL1B-indcued IRF1 mRNA at 1 h (Fig. 5.2B). However, this repressive effect of the DUSP1 adenovirus was lost by 2 h, and at 6 h, there was a significant enhancement (2.6 ± 0.4 fold) of IRF1 expression over that obtained with IL1B alone (Fig. 5.2B). Analysis of the effect of DUSP1 over-expression on unspliced nuclear

IRF1 using RNA from the experiments depicted in Fig. 5.1A revealed that expression of unspliced nuclear IRF1 RNA to be significantly enhanced following DUSP1 over-expression (Fig. 5.2C).

Thus, the DUSP1-mediated enhancement of IRF1 mRNA may involve enhanced IRF1 transcription.

133

A time (h) 1 2 4 6 Ad-GFP + + + + Ad-DUSP1 + + + + IL1B - + + + - + + + - + + + - + + + DUSP1 NS IRF1 GAPDH

1.0 *

0.5

IRF1

protein /GAPDH 0.0 B C time (h) 1 2 6 time (h) 6 Ad-GFP + + + + + + Ad-GFP + Ad-DUSP1 + + + + + + Ad-DUSP1 + IL1B - - - + + + - - - + + + - - - + + + IL1B - + + + 1.0 400 ***

0.5 200

*** *** RNA IRF1

mRNA *** ***

(%IL1B) /GAPDH 0.0 unIRF1/U6 0 Figure 5.2 Effect of DUSP1 over-expression on IL1B-induced IRF1 expression. A, A549 cells were either not infected or infected with Ad5-DUSP1, or Ad5-GFP at a MOI of 10 for 24 h before IL1B treatment (1 ng/ml). Cells were harvested after 1, 2, 4 or 6 h for western blot analysis of DUSP1, IRF1 and GAPDH. Representative blots are shown. Following densitometric analysis, data (N = 4) were normalized to GAPDH and plotted as means ± S.E. B, Cells were treated as in A and harvested at 1, 2 or 6 h for real-time PCR analysis of IRF1 and GAPDH. Data (N = 4) were normalized to GAPDH and plotted as means ± S.E. For A & B, significance relative to time- matched IL1B and/orAd5-GFP-treated samples was tested by ANOVA with a Bonferroni multiple comparison test. *, p < 0.05; ***, p< 0.001. C, Cells were treated as in A and harvested after 6 h for real-time PCR analysis of unspliced nuclear (un) IRF1 RNA and U6 RNA. Data (N = 4), normalized to U6, were expressed as a percentage of IL1B-stimulated cells and plotted as means ± S.E. Significance, relative to IL1B-treated samples was tested by ANOVA with a Dunnett's post- test.***, p< 0.001. NS = nonspecific band

5.3.5 Effect of MAPK inhibitors on IRF1 expression

Since DUSP1 over-expression attenuates IL1B-induced MAPK activation in A549 cells (chapter three), the effect of MAPK inhibition on IL1B-induced IRF1 was also tested. MAPK pathways

134

were inhibited using maximally effective concentrations of the p38 inhibitor, SB203580 (10 µM), the MAPK/ERK kinase (MEK) 1/2 inhibitor, U0126 (10 µM) or the JNK inhibitor, JNK inhibitor

8 (JNK-IN-8) (10 µM), individually and in combination to inhibit all three MAPK pathways simultaneously in the presence of IL1B (Fig. 5.3A). IL1B-induced IRF1 protein expression was not affected by the MAPK inhibitors either alone or in combination at 2 h. However, by 4 h post-

IL1B treatment, when IL1B-induced IRF1 protein expression was starting to decline, SB203580, or three inhibitors combined, maintained IRF1 protein expression to significantly higher levels compared to IL1B-induced IRF1 protein. At 6 h, IL1B-induced IRF1 protein had further declined, but SB203580 and the combination inhibitor treatment significantly attenuated this loss (Fig.

5.3A).

To explore the mechanistic basis of IRF1 regulation by MAPKs, A549 cells were treated with

IL1B for various times in the presence of the MAPK pathway inhibitors, as in Fig. 5.3A, and IRF1 mRNA expression was analyzed by qPCR. Following 1 h of IL1B treatment, IRF1 mRNA was rapidly induced, reached a maximum level at 2 h, and declined at 4 and 6 h (Fig. 5.3B, left panel).

The MAPK pathway inhibitors, alone or in combination, had no significant effect on IL1B-induced

IRF1 mRNA at 1 or 2, although the combination of MAPK inhibitors produced a modest attenuation at 1 h (Fig. 5.3B, right panel). By 4 h, IRF1 mRNA expression was significantly enhanced by SB203580 and the three inhibitors together relative to IL1B-treated and this effect was more pronounced at 6 h (Fig. 5.3B, right panel).

135

B A time (h) 2 4 6 JNK-IN-8 + + + + + + 300 SB203580 300 UO126 UO126 + + + + + + + + + + + + 1.5 200 ** 200 SB203580 NT ** IL1B - + + + + + - + + + + + - + + + + + *** 1.0 IL1B 100 100 IRF1 0 0 0.5

mRNA JNK-IN-8 SB+UO+J8 ** 300 300

GAPDH *** *** *** each time)each

IRF1/GAPDH 0.0 (% of IL1B at at IL1B (%of

IRF1/GAPDH 200 200 6 0 2 4 6 *** ** time (h) 100 100 4 ** * *** 0 0 1 2 4 6 1 2 4 6 protein 2 *** * time (h) time (h)

IRF1/GAPDH 0 C D 400 SB203580 400 UO126 300 30 min 150 90 min 300 300 1.5 *** 200 100 NT 200 200 * * * 50 1.0 *** IL1B 100 100 100 *** 0 0 0 0

RNA 0.5 400 400 JNK-IN-8 SB+UO+J8 150 120 min 150 180 min *** ** mRNA

unIRF1/U6 *** (% t = (%= t 0)

each time)each 300 300

0.0 unIRF1/U6 100 100 (% of IL1B at at IL1B (%of 0 2 4 6 200 200 * IRF1/GAPDH time (h) 100 100 50 50 0 0 0 0 1 2 4 6 1 2 4 6 0 15 30 45 0 15 30 45 time (h) time (h) time (h) time (h) E 300 IL1B treatment time (min) 30 90 120 180 240 IL1B 200 IL1B + SB203580 + + + + + IL1B + + + + + + + + + + SB203580 100 *

mRNA * *** (% t = (%= t 0) 200 50 50 50 50

IRF1/GAPDH 0

0 20 40 60 100 25 25 25 25 mRNA

time (min) (%= t 0)

IRF1/GAPDH 0 0 0 0 0 Figure 5.3 Effect of MAPK inhibitors on IRF1 expression. A, A549 cells were either not treated, treated with IL1B (1 ng/ml) or pre-treated with either UO126, SB203580, JNK inhibitor 8 (JNK-IN-8) or a combination of UO126, SB203580 plus JNK-IN-8 each at 10 µM for 30 min prior to IL1B stimulation. Cells were harvested at the indicated times prior to western blot analysis of IRF1 and GAPDH. Representative blots are shown. Following densitometric analysis, data (N = 4) were normalized to GAPDH and plotted as means ± S.E. B & C, Cells were treated as in A and harvested at 1, 2, 4 or 6 h for real-time PCR analysis of (B) IRF1 and GAPDH or (C) unspliced nuclear (un) IRF1 RNA and U6. Data (N = 4) were normalized to GAPDH or U6 and are plotted as means ± S.E. The effect of IL1B + MAPK inhibitors for IRF1 mRNA or unspliced nuclear (un) IRF1 RNA is plotted as a percentage of IL1B at the indicated times. For A, B & C, significance, relative to time-matched non-treated or IL1B-treated samples was tested using ANOVA with a Bonferroni multiple comparison test. *, p < 0.05; **, p < 0.01; ***, p < 0.001. D, A549 cells were treated with IL1B (1 ng/ml) for 30, 90, 120 and 180 min. Actinomycin D (Act D, 10 μg/ml) was then added (t = 0) and the cells were harvested as indicated. RNA was extracted for real-time PCR analysis of IRF1 and GAPDH. Data (N = 3) were normalized to GAPDH and are plotted as a percentage of t = 0 for each treatment as means ± S.E. E, A549 cells were treated with IL1B (1 ng/ml) or pre-treated for 30 min with SB203580 at 10 µM prior to IL1B stimulation for 30 min. Following this (t = 0), actinomycin D (Act D, 10 μg/ml) was added and cells were harvested at the indicated times. RNA was extracted for real-time PCR analysis of IRF1 and GAPDH. (Cont.…)

136

Figure 5.3 Continued. Data (N = 4) normalized to GAPDH, are plotted as a percentage of t = 0 for each treatment as means ± S.E. (left panel). Data following 45 min of Act D treatment are also shown (right panel). A549 cells were also treated with IL1B or IL1B + SB203580 at 10 µM for 90, 120, 180 and 240 min, prior to the addition of Act D (t = 0) for 45 min and analyzed as above (right panels). Significance, relative to time matched IL1B-treated samples was tested using a paired t-test. *, p < 0.05; ***, p < 0.001.

5.3.6 Effect of MAPK inhibitors on IRF1 transcription and mRNA stability

Analysis of unspliced nuclear IRF1 RNA revealed that the peak in transcription rate occurred at or before 1 h post-IL1B treatment, and then declined rapidly to reach near basal levels by 6 h (Fig.

5.3C, left panel). At 1 and 2 h, IRF1 transcription was high and the MAPK pathway inhibitors had no effect (Fig. 5.3C, right panel). However, the unspliced nuclear IRF1 RNA induced by IL1B was significantly enhanced at 4 h by SB203580, UO126 and all three inhibitors together (Fig.

5.3C, right panel). Similar, but lesser effects were observed at 6 h (Fig. 5.3C, right panel). Taken together, these data suggest that the induction phase of IRF1 mRNA expression was only marginally affected by MAPK inhibitors. However, inhibition of MAPKs reduced the rapid loss of IRF1 mRNA occurring after the peak in mRNA expression at 2 h. This effect involved a failure to attenuate IRF1 transcription at the later time points.

The stability of IRF1 mRNA following MAPK inhibition was also assessed using actinomycin D chase methodology. Following short, 30 min, of IL1B treatment, IRF1 mRNA appeared to be highly stable and did not decline below the 100% level over the 60 min of the actinomycin D chase period (Fig. 5.3D). Instead, IRF1 mRNA levels continued to rise immediately post-actinomycin D treatment. As suggested for TNF mRNA stability in chapter four, this may also be due to the fact that while transcription is blocked by actinomycin D, the prior accumulation of pre-formed unspliced/unprocessed-mRNA may allow the production of mature IRF1 mRNA to continue

137

briefly following actinomycin D treatment (378). However, this apparent rise in IRF1 mRNA was not observed at longer IL1B treatment times (90, 120 or 180 min post IL1B treatment) (Fig. 5.3D).

Rather, there was a loss of IRF1 mRNA to ~50% of starting levels (t = 0) over the course of the actinomycin D chase following longer (90, 120 and 180 min) IL1B treatment times (Fig. 5.3D).

Since the p38 MAPK was a major regulator of IL1B-induced IRF1 mRNA and protein expression

(see above), the effect of SB203580 was also examined on IRF1 mRNA stability. At 30 min, there was no loss of IL1B-induced IRF1 mRNA by SB203580 over the course of the actinomycin D chase (Fig. 5.3E, left panel). However, the continued increase in IRF1 mRNA post-actinomycin D treatment was inhibited by SB203580. This suggests that p38 MAPK may play a role in the post- transcriptional processing of unprocessed-RNA to mature IRF1 mRNA. This needs further investigation. Conversely, SB203580 had no effect on IRF1 mRNA stability following 90 min of

IL1B treatment, and at all later times (120, 180 and 240 min), IRF1 mRNA stability was modestly enhanced by SB203580, with significance being achieved only at 120 min post-treatment (Fig.

5.3E, right panels).

5.3.7 Effect of MAPK inhibitors on IRF1 protein stability

IRF1 protein is a short-lived protein with a t½ of around 30 min (408, 409). To determine the effect of MAPK inhibition on IRF1 protein stability, a chase methodology was adopted using the translational inhibitor, cycloheximide. Following 2 h of IL1B treatment, IL1B-induced IRF1 protein expression was profoundly inhibited by the addition of cycloheximide (Fig. 5.4A).

138

CHX A IL1B CHX B IL1B MG132 time (h) -2 0 ¼ ½ 1 time (h) -2 0 1

Harvest Harvest

time (h) 0 ¼ ½ 1 CHX + + CHX + + + MG132 + + IL1B - + + + + IL1B - + + + + IRF1 IRF1

GAPDH GAPDH 150 *** 300 * ** ** 100 200 * ***

50 IRF1

IRF1 100

protein

protein

/GAPDH

(%IL1B) /GAPDH 0 (%IL1B) 0

CHX C IL1B MG132 D SB+UO+J8 IL1B CHX time (h) -2 0 ¼ ½ 1 time (h) -2½ -2 0 ¼ ½ 1

Harvest Harvest

time (h) 0 ¼ ½ 1 0 ¼ ½ 1 time (h) 0 ¼ ½ 1 0 ¼ ½ 1 MG132 + + + SB/UO/J8 + + + + CHX + + + + + + CHX + + + + + + IL1B - + + + + + + + + IL1B - + + + + + + + + IRF1 IRF1 GAPDH GAPDH

100 + CHX 100 + CHX + MG132 + SB+UO +J8

50 * 50 *

protein (%IL1B) protein 0

IRF1/GAPDH 0 (% of control) (%of 0 20 40 60 IRF1/GAPDH 0 20 40 60 time (min) time (min) Figure 5.4 Effect of MAPK inhibitors on IRF1 protein stability. A, A549 cells were pre-treated with IL1B (1 ng/ml) for 2 h. Following this (t = 0), cycloheximide (CHX, 10 μg/ml) was added and cells were harvested at the indicated times for western blot analysis of IRF1 and GAPDH. Representative blots are shown. Following densitometric analysis, data (N = 4), normalized to GAPDH, were expressed as a percentage of IL1B-stimulated cells and plotted as means ± S.E. Significance relative to IL1B-treated samples, was tested by ANOVA with a Dunnett's post-test. **, p< 0.01; ***, p< 0.001. B, A549 cells were pre-treated with IL1B (1 ng/ml) for 2 h. Following this (t = 0), cycloheximide (CHX, 10 μg/ml) and/or MG132 (10 μg/ml) was added and cells were harvested after 1 h prior to western blot analysis of IRF1 and GAPDH. Representative blots are shown. Following densitometric analysis, data (N = 4), normalized to GAPDH, were (cont.…)

139

Figure 5.4 Continued. expressed as a percentage of IL1B-stimulated cells and plotted as means ± S.E. Significance, relative to either IL1B or IL1B+MG132-treated samples was tested using ANOVA with a Bonferroni multiple comparison test. *, p < 0.05; **, p < 0.01; ***, p < 0.001. C, Cells were treated as in B and harvested at the indicated times prior to western blot analysis of IRF1 and GAPDH. Representative blots are shown. Following densitometric analysis, data (N = 4), normalized to GAPDH were expressed as a percentage of IL1B and plotted as means ± S.E. Significance, relative to IL1B+CHX-treated samples was tested using a paired t-test. *, p < 0.05. D, A549 cells were either not treated, treated with IL1B (1 ng/ml) or pre-treated with a combination of UO126, SB203580 plus JNK inhibitor 8 (JNK-IN-8) each at 10 µM for 30 min prior to IL1B stimulation for 2 h. Following this (t = 0), cycloheximide (CHX, 10 μg/ml) was added and cells were harvested at the indicated times prior to western blot analysis of IRF1 and GAPDH. Representative blots are shown. Following densitometric analysis, data (N = 4), normalized to GAPDH were expressed as a percentage of IL1B or IL1B+MAPK inhibitors and plotted as means ± S.E. Significance, relative to IL1B+CHX-treated samples was tested using a paired t-test. *, p < 0.05.

Within 30 min of cycloheximide treatment IL1B-induced IRF1 was significantly decreased, and by 60 min post-cycloheximide treatment, IRF1 protein was almost completely lost (t½ = ~30 mins).

Conversely, IL1B-induced IRF1 protein expression was significantly enhanced by the proteasome inhibitor, MG132 (Fig. 5.4B). In the presence of cycloheximide, the stability of IRF1 protein was modestly increased by MG132 (Fig. 5.4B & C). Like MG132, combined MAPK inhibitors also produced a partial stabilisation of IL1B-induced IRF1 protein (Fig. 5.4D).

5.3.8 Characterization of IRF1 expression in the presence of IL1B and dexamethasone

Since MAPK activity is reduced by glucocorticoids in A549 cells (114, 283, 382), the effect of the synthetic glucocorticoid, dexamethasone, was examined on IL1B-induced IRF1 expression. IRF1 protein was rapidly induced by IL1B, reached a peak at 2 h, before declining at 4 and 6 h (Fig.

5.5A & B). IL1B-induced IRF1 protein was modestly (32.5 ± 5.6%), but significantly, repressed by co-treatment with dexamethasone at 2 h, but this repressive effect was reversed by 4 h, and at

6 h there was no repression of IRF1 protein by dexamethasone (Fig. 5.5A & B).

140

A B C 2 6 18 time (h) 1 time (h) 1 2 4 6 time (h) 1 2 6 Dex + + + + + + + + Dex + + + + Dex + + + + + + IL1B - - + + - - + + - - + + - - + + IL1B - + + + + + + + + IL1B - - + + - - + + - - + + NS IRF1 IRF1 IRF1 GAPDH GAPDH GAPDH 150 *** 150 *** *** ** * 0.6 100 100 0.4

50 *** protein

protein 50

*** IRF1 0.2

protein

(% IL1B 2 h) 2 (%IL1B

/GAPDH IRF1/GAPDH

0 h) 2 (%IL1B IRF1/GAPDH 0 0.0

IL1B D E G Dex CHX 45 1.5 *** 100 * NT * * 1.0 time (h) -4 0 ¼ ½ 1 30 * * Dex *** 50 IL1B

15 0.5 * mRNA

mRNA ** IL1B+Dex each time)each *** ** IL1B+Dex ** * Harvest (% of IL1B at at IL1B (%of 0.0 0 0 IRF1/GAPDH IRF1/GAPDH 0 ¼ ½ 1 0 ¼ ½ 1

(Fold of 1 h NT) h 1 of (Fold time (h) 0 2 4 6 1 2 4 6 0.5 1 2 6 18 time (h) CHX + + + + + + 1.0 *** 100 F Dex + + + + ** IL1B IL1B - + + + + + + + + 0.5 50 *** IL1B+Dex RNA IRF1

*** 150 each time)each

** IL1B+Dex unIRF1/U6

0.0 at IL1B (%of 0 100 GAPDH 0 2 4 6 1 2 4 6 time (h) time (h) 50 100 Control

mRNA

(% t = 0) +Dex 0

NT IL1B IRF1/GAPDH 0 20 40 60 50

time (min) * protein

0 (% of control) (%of IRF1/GAPDH 0 20 40 60 time (min)

Figure 5.5 Characterization of IRF1 expression in the presence of IL1B and dexamethasone. A&B, A549 cells were either not treated or stimulated with IL1B (1 ng/ml), dexamethasone (Dex, 1 μM) or a combination of the two as indicated. Cells were harvested at the indicated times prior to western blot analysis of IRF1 and GAPDH. Representative blots are shown. Following densitometric analysis, data (N = 6), normalized to GAPDH were expressed as a percentage of 2 h IL1B-stimulated cells and plotted as means ± S.E. For A, significance, using ANOVA with a Bonferroni's multiple comparison test is indicated. For B, significance relative to IL1B-treated samples was tested by ANOVA with a Dunnett's post-test.*, p < 0.05; **, p< 0.01; ***, p< 0.001. C, HBE cells were not treated (NT) or treated as in A and harvested at the indicated times for western blot analysis of IRF1 and GAPDH. Representative blots are shown. Following densitometric analysis, data (N = 4) were normalized to GAPDH and plotted as means ± S.E. Significance, using ANOVA with a Dunnett's post-test is indicated. ***, p< 0.001. D, A549 cells were treated as in A and harvested at the indicated times for real-time PCR analysis of IRF1 and GAPDH mRNA (upper left panel) or unspliced nuclear (un) IRF1 RNA and U6 RNA (lower left panel). Data (N = 4 - 11), normalized to GAPDH were expressed as a fold of 1 h NT (upper left panel) or were normalized to U6 (lower left panel), and plotted as means ± S.E. The effect of IL1B + dexamethasone for IRF1 mRNA (upper right panel) (N = 11) and unspliced nuclear (un) TNF RNA (lower right panel) (N = 4) is plotted as a percentage of IL1B for indicated times. Significance, relative to time matched NT or IL1B-treated samples was tested using a Bonferroni's multiple comparison test. *, p < 0.05; **, p< 0.01; ***, p< 0.001. E, HBE cells were (cont....)

141

Figure 5.5 Continued. treated as in A and harvested at the indicated times for real-time PCR analysis of IRF1 and GAPDH mRNA. Data (N = 4), normalized to GAPDH and plotted as means ± S.E. Significance, using ANOVA with a Bonferroni's multiple comparison test is indicated. *, p < 0.05; **, p< 0.01. F, A549 cells were treated with IL1B (1 ng/ml) or IL1B plus dexamethasone (Dex, 1 μM) for 90 min. Following this (t = 0), actinomycin D (Act D 10 μg/ml) was added and the cells were harvested at the indicated times for real-time PCR analysis of IRF1 and GAPDH. Data (N = 4) were normalized to GAPDH and plotted as a percentage of t = 0 for each treatment as means ± S.E. G, A549 cells were treated with IL1B (1 ng/ml) or IL1B plus dexamethasone (Dex, 1 μM) for 4 h. Following this (t = 0), cycloheximide (CHX, 10 μg/ml) was added and cells were harvested at the indicated times prior to western blot analysis of IRF1 and GAPDH. Representative blots are shown. Following densitometric analysis, data (N=4), normalized to GAPDH were expressed as a percentage of IL1B or IL1B+Dex and plotted as means ± S.E. Significance, relative to IL1B+CHX-treated samples was tested using a paired t-test. *, p < 0.05. NS = nonspecific band.

In primary HBE cells, IRF1 protein was not induced by IL1B at 1 h post-treatment. However, at 2 h, IRF1 protein expression was significantly increased by IL1B and was almost completely diminished by 6 h (Fig, 5.5C). There was no effect of dexamethasone on IRF1 protein expression.

In A549 cells, IRF1 mRNA was rapidly induced by IL1B at 1 h, and this continued to increase, reaching a peak at 2 h, before declining at 4 and 6 h (Fig. 5.5D, upper left). At each time, IRF1 mRNA was only 30 - 40% repressed by dexamethasone (Fig. 5.5D, upper right). In primary HBE cells, similar modest, but non-significant, repressive effects of dexamethasone were observed on

IL1B-induced IRF1 mRNA (Fig. 5.5E).

5.3.9 Effect of dexamethasone on IRF1 transcription rate

Analysis of unspliced nuclear RNA revealed that IL1B-induced IRF1 transcription was significantly (~50% loss) inhibited by dexamethasone at 1 h post-treatment (Fig. 5.5D, lower right). Since IL1B-induced IRF1 mRNA expression is blocked by the dominant inhibitor of NF-

κB, IκBα∆N, and repression of IRF1 mRNA by dexamethasone is unaffected by cycloheximide

142

(273), dexamethasone-mediated loss of IRF1 transcription could be attributed to classical NR3C1 transrepression. Indeed, TNF-induced IRF1 expression was only modestly repressed by dexamethasone in bronchial epithelial BEAS-2B cells (Shah S. unpublished data). Furthermore, binding of p65 (RELA) to a site upstream of IRF1 transcription start is also attenuated by dexamethasone (Dr. Anthony Gerber – personal communication). Thus, the effect of the maximum effective concentration of selective IKK2 inhibitor, PS1145 (10 M), on IRF1 expression was tested. IL1B-induced IRF1 mRNA was significantly, 71.4 ± 5.3 %, repressed by PS1145 at 2 h

(Fig. 5.6). This confirms a role for NF-κB in IRF1 expression and supports the concept of direct

NR3C1 transrepression in the dexamethasone-mediated repression of IRF1. Regardless, the repressive effect of dexamethasone on IL1B-induced unspliced IRF1 nuclear RNA was completely lost at 2, 4 and 6 h post-IL1B treatment (Fig. 5.5D, lower right). Thus, the net effect of dexamethasone on IRF1 mRNA expression was relatively minor.

time (h) 2 Figure 5.6 Effect of PS1145 on IL1B- PS1145 + induced IRF1 expression. A549 cells were IL1B - + + either not stimulated, treated with IL1B (1 150 IRF1 ng/ml) or pre-treated with PS1145 (10 µM) * 100 for 30 min prior to IL1B stimulation. Cells

50 were harvested after 2 h for real-time PCR mRNA (%IL1B) analysis of IRF1 and GAPDH. Data (N = 4),

IRF1/GAPDH 0 normalized to GAPDH were expressed as a percentage of IL1B-stimulated cells and plotted as means ± S.E. Significance, relative to IL1B-treated samples was tested using a

paired t-test. *, p < 0.05.

143

5.3.10 Effect of dexamethasone on IRF1 mRNA and protein stability

Following 90 min of IL1B treatment, IRF1 mRNA was reduced to ~50% of initial levels (t = 0) within 45 min of actinomycin D addition (Fig. 5.5F). This was unaffected by dexamethasone co- treatment. The effect of dexamethasone co-treatment on IRF1 protein stability was also analyzed.

IL1B-induced IRF1 protein was rapidly lost following the addition of cycloheximide. This loss of

IRF1 protein was modestly, but significantly, reduced in the presence of dexamethasone (Fig.

5.5G).

5.3.11 Effect of dexamethasone on IRF1-dependent gene expression

Expression of the IRF1-dependent IL1B-induced mRNAs identified in Fig. 5.1F, was analysed in the presence of dexamethasone at 6 h. Co-treatment with dexamethasone produced a variable repression of IL1B-induced 10 IRF1-dependent mRNAs (Fig. 5.7A). While CMPK2, IFIT1, MX1,

IFIT3 isoform 2 and HELZ2 were highly (>88 %) repressed, UBD, APOL6, CFB and IFIT3 isoform 1 were only partially (50 – 60 %) attenuated by dexamethasone (Fig. 5.7A). IL1B-induced

CXCL10 mRNA was unaffected by dexamethasone co-treatment (Fig. 5.7A).

In the case of protein expression, IL1B-indued CXCL10, released into the supernatants, was undetectable up to 4 h post-IL1B treatment. Following this, CXCL10 release was gradually enhanced by IL1B and reached a plateau at ~18 h (Fig. 5.7B). Co-treatment with dexamethasone for up to 8 h did not affect IL1B-induced CXCL10 release. Thereafter, dexamethasone produced a partial, but non-significant, repression of CXCL10 release (Fig. 5.7B).

144

A NS Dex IL1B IL1B+DexGene p B CXCL10 200 IFIT3iso1 * 150 APOL6 ** 100 CFB **

(pg/ml) 50 Release Release UBD * CXCL10 HELZ2 *** 0 IFIT3iso2 ** 0 5 10 15 20 MX1 * time (h) IFIT1 *** CMPK2 *** NT IL1B Dex IL1B+Dex

0 100 % IL1B

Figure 5.7 Effect of dexamethasone on IRF1-dependent gene expression. A, A549 cells were either not treated (NT) or stimulated with IL1B (1 ng/ml), dexamethasone (Dex, 1 μM) or a combination of the two as indicated. Cells were harvested at 6 h for real-time PCR analysis of indicated genes and GAPDH. Data (N = 4), normalized to GAPDH were expressed as a percentage of IL1B and plotted as means ± S.E. Significance, relative to IL1B-treated samples was tested using a paired t-test. *, p < 0.05; **, p< 0.01; ***, p< 0.001. B, Cells were treated as in A and the supernatants harvested at indicated times for CXCL10 release measurement. Data (N = 4) expressed in pg/ml are plotted as means ± S.E.

5.3.12 Effect of DUSP1 silencing on MAPKs and IRF1 expression

Since dexamethasone reduces MAPK activity by inducing the expression of DUSP1 (chapter three), the effect of DUSP1 knock-down on IRF1 protein expression was also examined. DUSP1 expression was induced by IL1B at 1 h and this was enhanced by co-treatment with dexamethasone

(Fig. 5.8A). As shown in chapter three, DUSP1-targeting siRNAs profoundly reduced DUSP1 expression and produced a marked increase in the phosphorylation of p38, ERK and JNK MAPKs in the presence of both IL1B and IL1B plus dexamethasone (Fig. 5.8A). LMNA siRNA was used as a control siRNA and had no effect on IL1B or IL1B plus dexamethasone-induced DUSP1 or on

MAPK activity (Fig. 5.8A). These data confirm that IL1B-induced MAPKs are negatively

145

regulated by DUSP1 and enhanced expression of DUSP1 by dexamethasone is involved in the repression of MAPKs by glucocorticoids at 1 h.

Western blot analysis for IRF1 was performed at 4 h, the time at which the loss of IL1B-induced

IRF1 was just starting to occur. IL1B-induced IRF1 protein expression was significantly reduced by the two DUSP1 siRNAs relative to LMNA control siRNA (Fig. 5.8B). The expression of IRF1 protein induced by IL1B plus dexamethasone was comparable to that with IL1B alone and is consistent with Fig. 5.5B. However, knock-down of DUSP1 produced a significant reduction in the expression of IRF1 in the presence of IL1B plus dexamethasone (Fig. 5.8B). These data support a role for DUSP1 in maintaining IRF1 expression in the presence of both IL1B and IL1B plus dexamethasone. A B time (h) 1 time (h) 4 DUSP1 siRNA 2 + + DUSP1 siRNA 2 + + DUSP1 siRNA 1 + + DUSP1 siRNA 1 + + LMNA siRNA + + LMNA siRNA + + Dex + + + + Dex + + + IL1B - + + + + + + + + IL1B - + + + + + + DUSP1 IRF1 P-p38 GAPDH P-ERK *** 150 * ** *

P-JNK 100 siRNA)

protein 50 GAPDH

IRF1/GAPDH 0 (%IL1B+LMNA C time (h) 6 6 Figure 5.8 Effect of DUSP1DUSP1-targeting siRNA siRNA 2 on+ IL1B + -induced MAPKs+ + and IRF1 expression. A&B, A549 cells were incubatedDUSP1 withsiRNA LMNA 1 + (control) + or DUSP1+ - specific + siRNAs. After 24 h, LMNA siRNA + + + + cells were treated with IL1B (1 ng/ml)Dex or IL1B plus+ + dexamethasone+ +(1 + μ +M) (Dex) as indicated. Cells were then harvested at 1 and 4 IL1Bh and- total+ + + proteins + + + were- prepared+ + + + +for + western blot analysis 200 200 of DUSP1, phospho-ERK (P-ERK), phospho-p38CXCL10 (P-p38), phosphoIFIT3iso2-JNK (P-JNK), IRF1 and * GAPDH. Blots representative of at least100 6 - 9 such* experiments100 are shown.* For B, following

*

mRNA

siRNA) (%IL1B densitometric analysis, data (N = 6),+LMNA normalized to GAPDH were expressed as a percentage of 0 0 LMNA siRNA plus IL1B-stimulatedGene/GAPDH cells and plotted as means ± S.E.Significance using ANOVA with a Bonferroni's multiple comparison test is indicated. Significance between: LMNA control siRNA plus IL1B and each of the DUSP1 targeting siRNAs plus IL1B, and the LMNA control plus IL1B plus Dex is shown. *, p < 0.05; **, p< 0.01; ***, p< 0.001.

146

5.3.13 Effect of DUSP1 silencing and MAPK inhibitors on IRF1-dependent gene expression

The effect of DUSP1 knock-down on IL1B-induced late-phase mRNAs was examined. IL1B- induced CXCL10 and IFIT3 isoform 2, which showed the greatest enhancements following

DUSP1 over-expression (Fig. 5.1A), both showed trends towards reduced IL1B-induced expression in the presence of the DUSP1 targeting siRNAs (Fig. 5.9A). In the case of IL1B plus dexamethasone, the loss of DUSP1, in the presence of both siRNAs, produced a modest, but significant, reduction in CXCL10 and IFIT3 isoform 2 mRNA expression (Fig. 5.9A). A similar, modest repressive effect, for at least one DUSP1 targeting siRNA, was observed for IFIT3 isoform

1 in the presence of IL1B plus dexamethasone (Fig. 5.9A). A number of the remaining late-phase mRNAs also showed trends towards reduced expression, but this did not reach significance.

Reasons for the variable effect of DUSP1 siRNAs on late-phase mRNA expression are multiple, but may include the fact that MAPKs have multiple, potentially opposing, roles in the regulation of late-phase gene expression. This is illustrated below.

The expression of IFIT3 isoform 1 and UBD were largely unaffected in the presence of MAPK inhibitors (Fig. 5.9B), whereas following the inhibition of all three MAPK pathways, IL1B- induced mRNA expression of CFB, CXCL10, HELZ2, IFIT1 and IFIT3 isoform 2 was significantly enhanced (Fig. 5.9B). Conversely, APOL6, CMPK2 and MX1 were attenuated by the combined MAPK inhibitors (Fig. 5.9B). IL1B-induced release of CXCL10 into the supernatant was significantly increased following MAPK inhibition (Fig. 5.9C).

147

A time (h) 6 6 DUSP1 siRNA 2 + + + + DUSP1 siRNA 1 + + + + LMNA siRNA + + + + Dex + + + + + + IL1B - + + + + + + - + + + + + + B C time (h) 18 200 200 time (h) 6 6 CFB APOL6 SB+UO+J8 + + SB+UO+J8 + IL1B - + + - + + IL1B - + + 100 100 150 APOL6 CFB * * 200 * 300 0 0 75 100 150

200 200 (pg/ml) Release Release CMPK2 CXCL10 CXCL10 0 0 0 * 100 100 * CMPK2 225 CXCL10 * 100 150 ** 0 0 75 200 200 HELZ2 IFIT1 0 0 200 HELZ2 IFIT1

100 100 ** 150 *** mRNA 100

75 Gene/GAPDH

0 0 mRNA (%IL1B)

200 200 Gene/GAPDH (%IL1B+LMNAsiRNA) 0 0 IFIT3iso1 IFIT3iso2 150 IFIT3 600 IFIT3 iso2 100 *** 100 * iso1 400 * * 75 200 0 0 0 0 200 200 MX1 UBD 150 MX1 150 UBD * 100 100 75 75

0 0 0 0

Figure 5.9 Effect of DUSP1-targeting siRNA and MAPK inhibitors on IRF1-dependent gene expression. A, A549 cells were incubated with LMNA (control) or DUSP1-specific siRNAs. After 24 h, cells were treated with IL1B (1 ng/ml) or IL1B plus dexamethasone (1 μM) (Dex) as indicated. Cells were then harvested at 6 h for real-time PCR analysis of indicated genes and GAPDH. Data (N = 6), normalized to GAPDH were expressed as a percentage of LMNA siRNA plus IL1B-stimulated cells and plotted as means ± S.E. Significance using ANOVA with a Bonferroni's multiple comparison test is indicated.*, p < 0.05; ***, p< 0.001. B, A549 cells were either not stimulated, treated with IL1B (1 ng/ml) or pre-treated with a combination of UO126, SB203580 plus JNK inhibitor 8 each at 10 µM (SB+UO+J8) for 30 min prior to IL1B stimulation. Cells were harvested after 6 h for real-time PCR analysis of the indicated genes and GAPDH. C, Cells were treated as in B and the supernatants harvested after 18 h for CXCL10 release measurement. Data (N = 4) expressed in pg/ml are plotted as means ± S.E. For B & C, significance, relative to IL1B-treated samples, was tested using a paired t-test. *, p < 0.05; **, p< 0.01; ***, p< 0.001.

5.3.14 Role of IRF1 in late-phase gene expression in the presence of IL1B and IL1B plus dexamethasone

A549 cells were treated with IL1B or IL1B plus dexamethasone in the presence of two independent

IRF1-targeting siRNAs. IRF1 expression was profoundly induced by IL1B at 2 h and this was

148

partially repressed by dexamethasone (Fig. 5.10A). In both cases, IRF1-targeting siRNAs substantially attenuated IRF1 expression (Fig. 5.10A). Likewise, the expression of the IL1B- induced late-phase mRNAs identified in Fig. 5.1F was also significantly reduced by IRF1 knock- down (Fig. 5.10B). In the context of dexamethasone, as shown in Fig. 5.7A, expression of these

10 late-phase mRNAs was variably affected by dexamethasone co-treatment. IL1B-induced

CXCL10 was unaffected and MX1, IFIT1, UBD or CMPK2 mRNAs were highly repressed by dexamethasone (Fig. 5.10B). However, regardless of the effect of dexamethasone on IL1B- induced late-phase mRNA expression, IRF1-targeting siRNAs produced a further attenuation of late-phase mRNA expression (Fig. 5.10B). This was significant in the case of APOL6, CFB,

CXCL10, HELZ2, IFIT3 isoform 1, MX1 and UBD. However, even though CMPK2, IFIT1 and

IFIT3 isoform 2 showed trends towards further reduction with IRF1-targeting siRNAs, they did not reach significance. Thus, IL1B-induced IRF1 appears necessary for late-phase mRNA expression in the presence of both IL1B and IL1B plus dexamethasone.

Since CXCL10 release was maximum at 18 h following IL1B treatment (Fig. 5.7B), 18 h time point was selected to analyze the effects of IRF1 knock-down. CXCL10 release was significantly induced by IL1B and this was not affected by dexamethasone (Fig. 5.10C, left panel). In both cases, the IRF1 targeting siRNAs produced a significant inhibition of CXCL10 release (Fig. 5.10C, left panel). Thus, IRF1 is required for CXCL10 production, even in the presence of dexamethasone. LMNA siRNA had no effect on IL1B or IL1B plus dexamethasone-induced

CXCL10 release (Fig. 5.10C, right panel).

149

A time (h) 2 IRF1 siRNA 2 + + IRF1 siRNA 1 + + LMNA siRNA + + Dex + + + IL1B - + + + + + + IRF-1 GAPDH B time (h) 6 6 6 6 6 IRF1 siRNA 2 + + + + + + + + + + IRF1 siRNA 1 + + + + + + + + + + LMNA siRNA + + + + + + + + + + Dex + + + + + + + + + + + + + + + IL1B - + + + + + + - + + + + + + - + + + + + + - + + + + + + - + + + + + + APOL6 CFB CMPK2 CXCL10 150 ** 150 ** 150 *** 150 ** 150 *** HELZ2 ** ** *** ** ** *** 100 ** 100 ** 100 100 ** 100 ** ** * 50 50 50 50 50 * 0 0 0 0 0 IFIT1 IFIT3 IFIT3 UBD 150 * 150 *** 150 ** 150 * MX1 150 **

mRNA * *** iso1 ** iso2 * *** 100 100 ** 100 100 100 ** ** Gene/GAPDH 50 50 50 50 ** 50 *** **

(%IL1B+LMNAsiRNA) 0 0 0 0 0 C time (h) 18 time (h) 18 IRF1 siRNA 2 + + LMNA siRNA + + IRF1 siRNA 1 + + Dex + + LMNA siRNA + + IL1B - + + + + Dex + + + IL1B - + + + + + + 15 20 * ** * 10 * ** *

10 (pg/ml) 5

Release Release

CXCL10

(pg/ml) Release Release CXCL10 0 0 Figure 5.10 Effect of IRF1-targeting siRNA on IL1B-induced inflammatory gene expression. A, A549 cells were incubated with LMNA (control) or IRF1-specific siRNAs. After 24 h, cells were treated with IL1B (1 ng/ml) or IL1B plus dexamethasone (1 μM) (Dex) as indicated. Cells were harvested at 2 h and total proteins were prepared for western blot analysis of IRF1 and GAPDH. Blots representative of at least 4 such experiments are shown. B, Cells were treated as in A and harvested at 6 h for real-time PCR analysis of indicated genes and GAPDH. Data (N = 4), normalized to GAPDH were expressed as a percentage of LMNA siRNA plus IL1B-stimulated cells and plotted as means ± S.E. C, Cells were treated as in A and the supernatants harvested after 18 h for CXCL10 release measurement. Data (N = 4) expressed in pg/ml are plotted as means ± S.E. Effect of LMNA siRNA on IL1B-induced CXCL10 release is shown in right panel. For B & C, significance was tested using ANOVA with a Bonferroni post-test. Significance between: LMNA control siRNA plus IL1B and each of the IRF1 targeting siRNAs plus IL1B, and the LMNA control plus IL1B plus Dex is shown. Other comparisons are specifically indicated.*, p < 0.05; **, p< 0.01; ***, p< 0.001.

150

5.3.15 Occupancy of IRF1 at the promoters of IRF1-dependent genes in the presence of IL1B and IL1B plus dexamethasone

To explore the mechanistic basis for the regulation of IL1B-induced late-phase gens, ChIP was used to analyse IRF1 occupancy at selected late-phase loci. For this analysis, IRF1 binding peaks, visualized in the UCSC genome browser and a bioinformatics software, MatInspector, were used to identify putative IRF1 binding regions within the CXCL10, IFIT1, IFIT3 and CMPK2 loci.

Occupancy for IRF1 was determined relative to non-occupied control regions (hMYOD1, hOLIG3, hMYOG) following IL1B and IL1B plus dexamethasone treatment for 4 h. IL1B treatment significantly increased IRF1 binding above baseline at tested regions associated with CXCL10 (-

254 to -172), IFIT1 (+180 to +320), IFIT3 (-16 to +79) and CMPK2 (-82 to +20). While co- treatment with dexamethasone produced no effect on IL1B-induced IRF1 occupancy at CXCL10 and IFIT3 loci, a modest, but not significant, reduction in IRF1 occupancy was observed in the case of IFIT1 and CMPK2 loci (Fig. 5.11). These data suggest that following the treatment of A549 cells with IL1B for 4 h, IRF1 was recruited to the late-phase gene promoters and this was not materially affected by dexamethasone co-treatment.

time (h) 4 4 Dex + + Figure 5.11 Characterization of IRF1 occupancy on IL1B - + + - + + inflammatory loci in the presence of IL1B and 15 15 * * dexamethasone. A549 cells were either not treated or 10 10 5 5 stimulated with IL1B (1 ng/ml) or IL1B plus 0 0 dexamethasone (1 μM) (Dex). Cells were harvested at 4 h CXCL10 IFIT1 for real-time PCR analysis of indicated and control genes. (-254 to -172) ( +180 to +320) (Fold) IRF1 occupancy was calculated as a difference between CT 30 * 30 *

IRF1occupancy 20 20 values for each target locus as compared with the geometric 10 10 mean of CT values of three control regions that are not 0 0 occupied by IRF1. Data (N = 4), normalized to control IFIT3 CMPK2 genes were expressed as a fold of non-treated cells and ( -16 to +79) ( -82 to +20) plotted as means ± S.E. Significance, relative to non- treated samples was tested using a Friedman test with Dunn’s multiple comparison test. *, p < 0.05.

151

5.4 Discussion

Repression of inflammatory gene expression is a major feature of glucocorticoid action and often involves inhibition of MAPK signalling processes via enhanced expression of DUSP1 (296).

However, in addition to exerting anti-inflammatory effects, glucocorticoids spare or enhance local innate immune and host defense responses of the epithelium (67, 410, 411). In chapter three,

DUSP1 over-expression was shown to repress IL1B-induced MAPK activation and a number of inflammatory mRNAs, including IL8, CSF2 and PTGS2. However, 19 out of 46 IL1B-induced inflammatory mRNAs, including the inflammatory transcription factor, IRF1, showed a significant enhancement following DUSP1 over-expression. Since DUSP1 over-expression inhibits MAPKs, the loss of MAPK activity by DUSP1 represents a valid explanation for the observed enhancement of IRF1 by DUSP1 (382). Thus, the role of MAPKs in the regulation of IRF1 and other inflammatory genes that were enhanced following DUSP1 over-expression was investigated using selective MAPK inhibitors. However, because inhibition of any single MAPK pathway may enhance the activity of other MAPK pathways via cross-feedback control mechanisms (361, 412,

413), and dexamethasone treatment inhibits all three MAPKs (382), MAPK inhibitors were used in combination.

IRF1 was induced by IL1B in A549 cells, and enhanced DUSP1 expression, by switching off

MAPKs, maintained IRF1 expression. Loss of MAPK activity resulted in enhanced IRF1 mRNA and protein expression, and was predominantly attributed to enhanced IRF1 transcription. In this context, MAPKs by their interaction with HATs and HDACs can affect the transcription (414).

Thus, ERK by binding to HDAC4 enhances its nuclear localization and promotes transcriptional

152

repression (415). In respect of IRF1 mRNA stability, even though SB203580 had no effect on mRNA stabilisation following 30 min of IL1B treatment, at longer IL1B treatment times,

SB203580 produced a partial stabilisation of IRF1 transcript. Thus, reducing MAPK activity promotes IRF1 mRNA expression. Additionally, IRF1 protein was a very unstable protein (t½ of

~30 mins) and subject to proteasomal degradation. This finding is in line with previous reports indicating that a number of short-lived transcription factors are degraded by the proteasome pathway (416-421). Furthermore, evidence also suggests that MAPK can also target the downstream protein for ubiquitination and degradation (422). Thus, as with the proteasomal inhibitor, MG132, IRF1 stability was also enhanced following the inhibition of MAPKs. However, in both cases, the effect was somewhat modest.

The loss of IL1B-induced DUSP1, at 1 h, in chapter three, was associated with enhanced phosphorylation of p38, ERK and JNK MAPKs. Since IRF1 mRNA and protein expression was reduced by MAPKs just after the peak of their expression (i.e. at 2), increased phosphorylated

MAPK expression in DUSP1 inhibited cells may result in reduced IRF1 expression at later time points. Indeed, in this chapter, silencing of IL1B-induced DUSP1 enhanced the phosphorylation of MAPKs at 1 h and consequently reduced IRF1 protein expression at 4 h post-IL1B treatment.

This confirms a regulatory network, whereby DUSP1 switches off MAPKs to maintain IL1B- induced IRF1 expression, and such data are consistent with the negative regulation of IRF1 by

MAPKs in DUSP1-/- mouse macrophages, primary HBE and BEAS-2B cells (128, 129, 405).

153

Kinetic analysis of IL1B-induced, DUSP1-enhanced, mRNAs revealed that except for ICAM1 and

STAT5A, the remaining 11 mRNAs (APOL6, CFB, CMPK2, CXCL10, HELZ2, IFIT1, IFIT3 isoform 1&2, MX1, TLR2 and UBD) were all maximally induced at 6-18 h. Indeed, the IL1B- induced late-phase mRNAs, with the exception of TLR2, were IRF1-dependent. This was consistent with the recruitment of IRF1 to the promoters of CMPK2, CXCL10, IFIT1 and IFIT3 following IL1B treatment. Furthermore, a role for IRF1 in the up-regulation of CFB and CXCL10 is also suggested and ChIP-seq data also show the binding of IRF1 at the APOL6, IFIT1 and IFIT3 genes (188, 423, 424). Thus, a positive role for IRF1 in late-phase gene expression was confirmed.

Furthermore, these data also explain the ability of DUSP1 to enhance the expression of these mRNAs via the maintenance of IRF1 expression. Because IL1B-induced TLR2 was not IRF1- dependent, enhancement of TLR2 by DUSP1 must occur via alternative mechanisms. Conversely,

CCL5 and CXCL5 mRNAs were enhanced following IRF1 knock-down. Thus, an inhibitory role for IRF1 in the regulation of CCL5 and CXCL5 is predicted. These data, with respect to the negative regulation of CCL5 by IRF1, are consistent with previous findings in astrocytoma cells

(425). Since CCL5 plays an important role in the pathogenesis of asthma, maintenance of IRF1 by

DUSP1 could be beneficial in such condition (426).

Even though IL1B-induced IRF1 expression was maintained by both combined MAPK inhibitors and DUSP1 over-expression, effects on downstream IRF1-dependent mRNAs were more variable.

Thus, expression of the 10 IL1B-induced mRNAs shown in Fig. 5.1F was enhanced following

DUSP1 over-expression. However, only 5 of the late-phase mRNAs, including CFB, CXCL10,

HELZ2, IFIT1 and IFIT3 isoform 2, showed enhanced expression following combined MAPK

154

inhibitor treatment, and the others were simply not affected, or reduced modestly. Explanations for this variable effects are multiple, but are likely to involve: i) the fact that the expression of late- phase genes may also be regulated by additional non MAPK pathways, and/or factors; and ii) the effect of simultaneous inhibition of MAPKs by combined use of small molecule inhibitors is not entirely identical with DUSP1 over-expression. In addition, the expression of late-phase genes by

IL1B was NF-κB dependent (Fig. 5.12), and MAPKs can show opposing effects on NF-κB- dependent gene expression (427, 428). For example, ERK pathway negatively regulates NF-B- dependent gene expression (428), whereas p38 MAPK may have positive effects (427). Equally, while the selectivity of these small molecule kinase inhibitors is generally considered good, a number of off-target effects are also well established (357, 429, 430). In addition, when DUSP1 is over-expressed to inactivate its preferred MAPK substrate, the activity of other MAPK family members may also get repressed/affected and have additional uncharacterised effects on late-phase gene expression (100, 361). Regardless of the mechanisms enhancing CXCL10 and IFIT3 isoform

2 following MAPK inhibition, DUSP1 switches off MAPK and enhances these mRNAs. Thus, following DUSP1 knock-down, CXCL10 and IFIT3 isoform 2 mRNAs showed trends towards reduced expression at 6 h. However, other late-phase genes were without any obvious effect and this suggests that, unlike IRF1, the regulation of late-phase gene expression by MAPKs may involve multiple, potentially opposing effects.

155

time (h) 6 6 6 6 6 PS1145 + + + + + IL1B - + + - + + - + + - + + - + + 150 APOL6 150 CFB 150 CMPK2 150 CXCL10 150 HELZ2 *** *** 100 *** 100 100 ** 100 ** 75 50 50 50 50 0 0 0 0 0

mRNA 150 IFIT1 200 IFIT3 200 IFIT3 150 MX1 150 UBD (%IL1B) * iso1 iso2 ** ** Gene/GAPDH 100 *** 100 *** 100 75 75 50 0 0 0 0 0

Figure 5.12 Effect of PS1145 on IL1B-induced IRF1-dependent inflammatory gene mRNA expression. A549 cells were either not stimulated, treated with IL1B (1 ng/ml) or pre-treated with PS1145 (10 µM) for 30 min prior to IL1B stimulation. Cells were harvested after 6 h for real-time PCR analysis of the indicated genes and GAPDH. Data (N = 4), normalized to GAPDH were expressed as a percentage of IL1B-stimulated cells and plotted as means ± S.E. Significance, relative to IL1B-treated samples was tested using a paired t-test. *, p < 0.05; **, p< 0.01; ***, p< 0.001.

Dexamethasone, in both A549 and primary HBE cells, in conjunction with IL1B produces greatest levels of DUSP1 expression (chapter three & four). In addition, as described in chapter three, enhanced DUSP1 expression, in A549 cells, was important for early onset repression of MAPKs by dexamethasone. Thus, even though co-treatment with dexamethasone produced 32.5 ± 5.6% repression of IRF1 protein at early time (i.e. 2 h), this repressive effect of dexamethasone disappeared at longer treatment times (i.e. 4 and 6 h), and could be attributed to enhanced MAPK repression following IL1B plus dexamethasone co-treatment at 4 or 6 h (283, 382). This was further confirmed by the fact that enhanced phosphorylation of MAPK, following the loss of IL1B plus dexamethasone-induced DUSP1, at 1 h, produced a significant loss of IRF1 protein at 4 h.

Furthermore, dexamethasone also produced a ~50% loss of IL1B-induced IRF1 transcription at 1 h. This effect was largely reflected in IRF1 mRNA levels, which showed ~25-30% repression by

156

dexamethasone at all times. Additionally, IL1B-induced IRF1 expression was NF-κB-dependent

(Fig. 5.6) and the repression of IRF1 mRNA by dexamethasone occurs in the presence of cycloheximide (51). Thus, direct NR3C1-mediated repression, i.e. transrepression, may account for this early effect on IRF1 transcription rate by dexamethasone. This is also supported by the data from BEAS-2B cells where TNF-induced IRF1 expression was reduced by dexamethasone

(Shah S. unpublished data) in a manner that correlated with reduced binding of p65 (RELA) to the

IRF1 promoter region (Dr. Anthony Gerber – personal communication). Therefore, the observed repression of IRF1 transcription by dexamethasone could be possibly due to classical NR3C1 transrepression. Regardless of the mechanisms repressing IRF1 transcription and mRNA, dexamethasone induces DUSP1 expression and this inhibits MAPK activation (24). Since MAPKs negatively regulate IRF1 expression (i.e. at 4 or 6 h), inhibition of MAPKs by dexamethasone may contribute towards restoring or maintaining IRF1 expression at longer times following IL1B plus dexamethasone treatment. Thus, the net repressive effect of dexamethasone on IRF1 mRNA was very modest. Furthermore, IRF1 mRNA stability was also not affected by dexamethasone. In addition, as with MAPK inhibition, dexamethasone also modestly enhanced the stability of IRF1 protein, further suggesting that enhanced protein stability may account for the maintenance of

IRF1 by dexamethasone.

In the case of late-phase genes, while IL1B-induced CXCL10 expression was insensitive to repression by dexamethasone, other late-phase genes showed a varying degree of repression by dexamethasone. This supports the previously described variable repressive effect of glucocorticoids on some of these genes (411, 431, 432). Furthermore, silencing of IL1B plus

157

dexamethasone-induced DUSP1 significantly attenuated the expression of CXCL10 and IFIT3 isoform 2. Since, both CXCL10 and IFIT3 isoform 2 were IRF1-dependent and the recruitment of

IRF1 at their promoters was not altered (at 4 h) by dexamethasone, these data confirm that glucocorticoids enhance DUSP1 expression to reduce MAPK signalling and thereby maintain

IRF1 and downstream gene expression.

These results in A549 cells suggest that in addition to producing anti-inflammatory actions, glucocorticoids may spare or enhance innate immune and host defense responses through DUSP1- mediated maintenance of IRF1 and CXCL10. In this respect, roles for IRF1 and CXCL10 in innate immunity, anti-viral defenses and in the development of TH1 immunity are suggested and may be promoted by DUSP1 (170, 304, 433). Conversely, enhanced IRF1 expression is associated with reduced glucocorticoid sensitivity (210, 407). Equally, a role for CXCL10 in airway inflammation, airway hyper-responsiveness and virus-induced asthma exacerbation is also indicated (130, 434).

Thus, maintenance of IRF1 and/or CXCL10 by DUSP1 may not be therapeutically desirable.

Furthermore, results presented here add to, indeed modify, the emerging concept of glucocorticoid- insensitivity and suggest that pathways involving maintenance of IRF1 could contribute to poor response to glucocorticoids. Indeed, contrary to initial expectations (296, 392), it may be that novel

NR3C1 ligands that have reduced ability to induce DUSP1 and/or inhibit MAPK pathways could show an improved efficacy in the context of IRF1-dependent inflammatory responses.

Finally, confirmation of some of the findings, with respect to the effects of IL1B and dexamethasone on DUSP1 (in chapter four) and IRF1 expression in primary HBE cells suggest

158

that the data presented herein are likely to be physiologically and therapeutically relevant. In addition, data from DUSP1 knockout mouse models also support these findings (128, 405). In summary, loss of glucocorticoid-induced DUSP1 may enhance glucocorticoid sensitivity in the context of certain host defense responses where IRF1 is a key driver. Thus, only by considering such complex interactions the existing anti-inflammatory glucocorticoid therapies can be improved.

159

Chapter 6 : Discussion

6.1 Feedback and feed-forward control of inflammatory gene expression

Appropriate regulation of inflammatory gene expression is central to inflammation and its resolution. Equally, understanding these processes will aid the identification of therapeutic agents that target inflammation. However, while the regulation of signal transduction and gene expression involves numerous signalling molecules leading to transcriptional, post-transcriptional, translational and often post-translational control of gene expression, it is routine to depict these assemblies as simple linear pathways. In this regard, the three major MAPK, p38, ERK and JNK, play an important role in the initiation and execution of inflammatory responses in response to inflammatory stimulus, and thus, have been considered as promising targets for the treatment of inflammatory diseases (362, 435). As such, phosphorylation and dephosphorylation of MAPKs and their downstream targets serve as molecular switches for modulating inflammatory processes and in fact the level and magnitude of each of these cellular targets is a highly regulated process

(436). In this context, MAPK pathways control the expression of inflammatory mediators via the phosphorylation of RNA-binding proteins, including ZFP36, and the activation of transcription factors, such as c-Jun, ATF-2, p53, Elk-1, NFAT, NF-B and/or AP-1 (94, 95, 97, 101, 102). Thus, several kinase inhibitors that block the activity of MAPKs either directly or indirectly via the inhibition of upstream/downstream modulators are being evaluated currently to be used as therapeutic agents for the treatment of a number of chronic inflammatory diseases, including asthma (71, 437). Conversely, MAPK-mediated regulation of phosphatases, mainly DUSPs, by limiting the activity of MAPKs, via dephosphorylation, limit inflammatory responses (438). In

160

addition, mathematical modelling of MAPK pathways also suggests that negative feedback control of MAPKs by DUSPs is crucial to control the kinetics of MAPK activity (439-441). Furthermore,

MK2-mediated phosphorylation of ZFP36 prevents the mRNA destabilising activity and function of ZFP36 and allows dynamic control of the expression of ARE-containing inflammatory transcripts, including cytokines and chemokines (309, 311, 442, 443). However, dephosphorylation of ZFP36, by phosphatases, such as PP2A, by enhancing the activity of ZFP36, promotes mRNA decay of pro-inflammatory transcripts (312). Thus, even though MAPK- dependent regulation of gene expression is often depicted to follow a linear cascade, in reality, it behaves in a non-linear fashion. This nonlinearity in MAPK signalling and effector responses can be attributed to MAPK-dependent regulation of feedback and feed-forward regulatory proteins, such as DUSP1 and ZFP36 respectively (See Figure 4.1). As a result, to understand the regulation of inflammatory gene expression by glucocorticoids, it is fundamental to consider the feedback and feed-forward regulatory processes that are induced by inflammatory stimuli. These are addressed, in detail, in this thesis.

In A549 cells, the expression of both DUSP1 and ZFP36 was induced by IL1B in a MAPK- dependent manner. In addition, in the presence of IL1B, DUSP1 provided a transient feedback control of MAPKs and multiple IL1B-induced ARE-containing inflammatory transcripts, including CCL2, CCL20, CSF2, CXCL1, CXCL2, CXCL3, IL6, IL8 and PTGS2. In contrast,

IL1B-induced ZFP36 expression was essential to limit the expression of IL1B-induced ARE- containing transcripts, such as TNF. Additionally, siRNA-mediated loss of DUSP1 not only enhanced MAPK phosphorylation (at 1 h) (chapter three), but also increased ZFP36 protein

161

expression (at 2 h) post-IL1B treatment (chapter four). Furthermore, even though the mRNA expression of a number of IL1B-induced ARE-containing transcripts (see above) was enhanced at

1 h by the prior loss of DUSP1, by 6 h post-IL1B treatment, loss of DUSP1 attenuated the mRNA expression of ARE-containing transcripts, including CCL2, CXCL3 and PTGS2 (chapter three).

This could be mediated via the increased expression of ZFP36 following the loss of DUSP1, and was confirmed using TNF as a model ARE-containing ZFP36 target gene (chapter four). Thus, in

A549 cells, ZFP36 was induced by IL1B to reduce TNF mRNA stability to exert feed-forward inhibition of TNF mRNA and protein expression. In addition, the attenuation of TNF mRNA at 6 h in the presence of DUSP1 siRNA was prevented by the additional loss of ZFP36. This effect was associated with reduced TNF mRNA stability, which was also inhibited by ZFP36 knock-down.

Thus, even though the current data is restricted to the feed-forward inhibition of TNF by ZFP36, the overall concept may also hold true for other ARE-containing inflammatory transcripts, such as

PTGS2, CSF2 and/or IL8, that are also potential targets of ZFP36 (215, 222, 313).

In conclusion, these data directly confirm that the loss of DUSP1 enhances ZFP36 expression to increase negative feed-forward control of ARE-containing inflammatory transcripts, such as TNF.

However, despite these data, loss of ZFP36 did not completely ablate all the effects of DUSP1 knock-down on TNF mRNA expression. This strongly suggests that there are other non-ZFP36- dependent feed-forward repressive effects that are also dependent on MAPKs and which may also impact inflammatory gene expression. Additionally, confirmation of the regulation of ZFP36 by

DUSP1 in primary human airway smooth muscle cells and mouse models suggest that the regulatory mechanisms described herein using the A549 cells are physiologically and

162

therapeutically relevant (320, 404, 405). In particular, Smallie et al., using DUSP1-/- mice, demonstrated that DUSP1 regulates the expression of a number of inflammatory mediators, including TNF, CXCL1, CXCL2 and IL1B, by controlling the activity of ZFP36 (405). Similarly, in airway smooth muscle cells, the mRNA destabilising activity of ZFP36 is controlled by p38

MAPK-mediated phosphorylation (404). Thus, DUSP1, by dephosphorylating p38 MAPK, increases the activity of ZFP36 to repress inflammatory gene expression (404). Taken together, the interaction between DUSP1- and ZFP36-mediated feedback and feed-forward mechanisms respectively reveals why modulation, or targeting, of a single component in a signalling cascade can often produce opposing effects on inflammatory gene expression. Therefore, these data emphasize the need to accurately model network behaviour in order to predict correct biological and/or appropriate therapeutic outcomes.

6.2 DUSP1-mediated regulation of inflammatory gene expression

MAPKs, mainly p38 and JNK, play an important role in the regulation of innate immune responses

(101, 353). In particular, the p38 MAPK pathway plays a crucial role in the development of TH1 immunity and in the production of IFN(444). Inadequate up-regulation of innate immune responses, through the enhanced production of cytokines and chemokines, may lead to the development of chronic inflammatory diseases, including asthma (294, 445). However, termination of MAPK signalling via negative feedback processes could limit the innate immune responses (294, 353, 446). In this regard, as mentioned above, enhanced DUSP1 expression plays an important role in the feedback inhibition of MAPKs. Thus, over-expression of DUSP1 resulted in loss of IL1B-induced MAPK activity and expression of 11 inflammatory cytokines and

163

chemokines (chapter three). These data are consistent with findings in DUSP1-/- mouse macrophages, where a number of cytokines and chemokines showed enhanced expression following LPS challenge (352, 353) .

However, contrary to the current dogma, 14 out of 46 IL1B-induced inflammatory mRNAs, including the interferon regulatory factor, IRF1, and the chemokine, CXCL10, were significantly enhanced following DUSP1 over-expression. In addition, 11 of the 14 DUSP1-enhanced mRNAs, including CXCL10, were induced by IL1B in an IRF1-dependent manner with a peak of their mRNA expression at either 6 or 18 h. Like DUSP1 over-expression, inhibition of MAPKs also prolonged IRF1 expression, an effect involving elevated transcription and increased mRNA and protein stability (chapter five). Conversely, DUSP1 silencing increased IL1B-induced MAPK phosphorylation (see above), while significantly reducing IRF1 protein expression (Fig. 6.1 A&B).

Consistent with this, a study from Korhonen et al. also showed that IRF1 expression was reduced in DUSP1-/- macrophages, and pharmacological inhibition of p38 MAPK augmented LPS-induced

IRF1 expression (128). Furthermore, inhibition of MEK1, in human airway epithelial cells, enhances the expression and the activity of HRV-induced IRF1 (129, 188). Thus, these data collectively confirm a regulatory network whereby inhibition of MAPKs maintains IRF1 expression.

Although IL1B-induced CCL5 and CXCL5 mRNAs were reduced significantly following DUSP1 over-expression, siRNA-mediated loss of IRF1 substantially enhanced the mRNA expression of

IL1B-induced CCL5 and CXCL5. This suggests that these mRNAs are negatively regulated by

164

IRF1 and is consistent with previous findings in astrocytoma cells, where a negative role for IRF1 on CCL5 expression has been reported (425). Since CCL5 may play an important role in the pathogenesis of asthma, by increasing the concentration of eosinophils in the airways, the maintenance of IRF1 expression by DUSP1 could be beneficial in asthma (10).

Taken together, these results suggest that besides the inhibition of MAPKs and inflammatory gene expression, DUSP1 may also spare or even enhance innate immune and host defense responses through the maintenance of IRF1 and CXCL10. In this respect, the role of DUSP1, IRF1 and

CXCL10 in innate immunity, anti-viral host defense and development of TH1 immunity is clearly indicated (128, 170, 195, 304, 433, 447). Furthermore, in mouse macrophages, DUSP1 deficiency resulted in reduced production of IRF1 and IL12, an important cytokine in the development of TH1 responses (128). Since IL12 antagonizes TH2 responses and inhibits allergic inflammation (25, 26,

448), maintenance of IRF1 by DUSP1 could be beneficial in such conditions.

Another potentially important finding of this thesis is that the mRNA expression of IL1B-induced

TLR2 was also enhanced by DUSP1 over-expression. Even though the time course analyses revealed that TLR2 was a late-phase gene (i.e. expression at 6 or 18 h), unlike other late-phase genes in this study (chapter five), TLR2 expression was not IRF1 dependent. Thus, IL1B-induced

TLR2 expression may involve mechanisms that are independent of IRF1. In this regard, IL1B- induced TLR2 expression, in HBE cells, is NF-B-dependent, and negatively regulated by p38 and JNK MAPKs (449). Thus, loss of MAPK activity, mediated by DUSP1 over-expression, may enhance IL1B-induced TLR2 expression.

165

In terms of glucocorticoids, by acting on airway epithelium, they suppress the production of epithelial-derived inflammatory mediators and inhibit the recruitment of inflammatory cells into the airways, resulting in reduced inflammatory responses (9). However, glucocorticoids also enhance the expression of scavenger receptors, TLRs, collectins, complement family members and other antimicrobial and antiviral proteins that are involved in host defense (450, 451). Moreover, glucocorticoids also spare or increase the expression of epithelial genes, involved in innate immunity, and thereby contribute towards the maintenance of epithelial innate immune responses

(67). In this regard, glucocorticoids enhance epithelial expression of TLR2 and a number of other

TLR3-mediated genes associated with host defense mechanisms (411, 452-455). Thus, glucocorticoids can also induce many innate immune responses and presumably this may lead to an enhancement of innate immunity. The data presented in this thesis clearly demonstrate that

DUSP1 expression was induced following dexamethasone treatment and was further enhanced by

IL1B co-treatment. Indeed, in chapter five, siRNA-mediated loss of IL1B plus dexamethasone- induced DUSP1 resulted in enhanced loss of IRF1 and CXCL10 (Fig. 6.1C). Thus, in addition to promoting anti-inflammatory actions of glucocorticoids, DUSP1 may play an important role in the maintenance of innate immune and host defense responses via the maintenance of IRF1 and

CXCL10 in the presence of glucocorticoids.

166

A B C IL1B treatment IL1B treatment + DUSP1 siRNA IL1B + glucocorticoid IL1B IL1B IL1B

DUSP1 MAPK IRF1 MAPK IRF1 GC DUSP1 MAPK IRF1

Late-phase Late-phase Other GC inflammatory inflammatory effectors gene gene expression expression

Figure 6.1 Regulation of IRF1 and IRF1-dependent late-phase gene expression by IL1B following DUSP1 silencing and glucocorticoid treatment. Schematics representing possible regulatory networks are shown along with summarized data from chapter five. A, IL1B treatment results in the activation of MAPKs. This, along with the activation of other signalling pathway and inflammatory transcription factors, e.g. NF-κB and AP-1 (not shown), enhances expression of the negative feed-back regulator, DUSP1. Following MAPK activation, the increased expression of DUSP1 is one mechanism by which MAPK activity is restored to basal. Expression of late-phase inflammatory genes by IL1B depends on IRF1 activation. IL1B-mediated activation of MAPKs negatively regulate the expression of IL1B-induced IRF1 and selected IRF1-dependent late-phase genes (e.g. CXCL10). B, With reduced DUSP1 expression there is reduced negative feedback control of MAPKs leading to enhanced MAPK activity (at 1 h) (bold) at the times where DUSP1 expression would have been elevated. This reduces IRF1 (grey) expression. C, In the presence of glucocorticoid co-treatment, DUSP1 expression is enhanced (bold) and promotes inactivation of MAPKs (grey). Loss of MAPK activity maintains the expression of IRF1 and selected IRF1- dependent late-phase genes (e.g. CXCL10) in the presence of glucocorticoids. However, since the expression of some of the IRF1-dependent late-phase genes (e.g. APOL6, CMPK2, CFB, HELZ2, IFIT1, IFIT3 iso1&2, MX1 and UBD) is variably, but significantly repressed by glucocorticoids, additional glucocorticoid effectors must exist to maintain repression of selected IRF1-dependent late-phase genes. Positive signalling/expression (blue) is represented by arrows. Negative effects are indicated (red) by lines ending in a T-bar.

6.3 Glucocorticoid-inducible genes and redundancy

Glucocorticoids repress inflammatory gene expression via mechanisms that may include direct repression of transcription (transrepression) and/or the induction of glucocorticoid-induced effector genes (transactivation) that then reduce inflammatory gene expression (456, 457).

Importantly, inflammatory gene expression may often involve MAPK activation and there is a

167

considerable interest in DUSP1 as a major effector of glucocorticoid repression (458, 459). Thus, the role of DUSP1 in glucocorticoid-induced inflammatory gene repression was explored using 11

IL1B-induced inflammatory genes whose mRNA expression was repressed by dexamethasone in a cycloheximide-sensitive manner (460). This implies that glucocorticoid-induced gene expression, i.e. transactivation, is crucial for repression. In this context, DUSP1 is strongly induced by dexamethasone in A549 cells and over-expression of DUSP1 produced a complete loss of

IL1B-induced MAPK activation and inflammatory gene expression (chapter three). Despite this, efficient knock-down of IL1B plus dexamethasone-induced DUSP1 did not completely reverse the repression of MAPKs and inflammatory genes by dexamethasone. In fact, only a transient, non- redundant role, for dexamethasone-induced DUSP1 in the early onset of repression of MAPKs and only certain IL1B-induced inflammatory genes (such as CXCL1, CXCL2 and PTGS2) tested here was documented (chapter three).

Explanations for this partial effect of DUSP1 loss could be many, and may involve DUSP1- independent effector mechanisms that may also be enhanced/induced by dexamethasone. For example, glucocorticoid-induced TSC22D3 and CDKN1C, which may inhibit Ras-Raf activation of MAPKs and JNK signaling respectively, are both induced by dexamethasone in A549 cells, and may therefore play an important role in dexamethasone-induced repression of MAPKs (282, 350,

363, 370-373). Moreover, expression of the IL1B-induced inflammatory genes studied in chapter three was mainly NF-B dependent (273). Thus, inhibition of NF-B-dependent transcription by both DUSP1 and TSC22D3, in addition to NFKBIA, which is modestly induced by dexamethasone in A549, primary HBE and airway smooth muscle cells (Leigh et al. unpublished data), may also

168

play a repressive role (114, 461-463). Additionally, NF-B-dependent transcription can also be inhibited by the transcriptional repressor, zinc finger and BTB domain containing 16 (ZBTB16)

(464). Recent data from our laboratory suggest that ZBTB16 mRNA is strongly induced by dexamethasone and budesonide in A549, primary HBE and airway smooth muscle cells (Leigh et al. unpublished data). Additionally, in A549 cells, ZBTB16 mRNA and protein was synergistically induced by IL1B plus dexamethasone co-treatment (Shah S. unpublished data).

Moreover, recent work by Sadler et al., using ZBTB16 knockout mice, demonstrated that ZBTB16 plays an important role in the attenuation of a number of NF-B-induced inflammatory gene expression (465). Thus, enhanced expression of ZBTB16 by glucocorticoids could also inhibit NF-

B and NF-B-dependent gene expression independent of DUSP1. Likewise, ZFP36 is also suggested to inhibit NF-B-dependent transcription by interacting with p65 subunit of NF-B and

HDAC1, -3, and -7 (466). In addition, TNFAIP3 (A20), a key inhibitor of NF-κB, is also induced by dexamethasone, and a feedback inhibitory role for dexamethasone-induced TNFAIP3 in NF-

κB inhibition is also established (467). Equally, IRAK-M (also known as IRAK3), a negative feedback regulator of the TLR/IL1R family signalling, is also induced by dexamethasone in A549 and BEAS-2B cells (468). Thus, even if the expression of DUSP1 was prevented, enhanced expression of TSC22D3, NFKBIA, ZBTB16, TNFAPI3, IRAK-M, ZFP36 and/or other anti- inflammatory effector genes (see section 1.3.4 for more details) by dexamethasone may still inhibit

NF-B and other inflammatory signaling processes.

The majority of inflammatory mRNAs studied in chapter three contain one or more AREs in their

3’-UTR region, and thus, the stability of such mRNAs could be inhibited by RNA binding proteins,

169

such as ZFP36. In this context, in chapter four, ZFP36 expression was shown to be modestly enhanced following dexamethasone treatment at 6 h in A549 cells, and is consistent with previous observation within our laboratory (308). Additionally, the repression of IL8 and CSF2 in A549 cells by dexamethasone was partly mediated by ZFP36 (308). Thus, even if DUSP1- and

TSC22D3-mediated repressive effects were prevented, ZFP36 may still exert repressive effects on

ARE-containing inflammatory gene expression at 6 h in the presence of dexamethasone. However, as confirmed in chapter four, siRNA-mediated loss of ZFP36 was not sufficient to reverse the repressive effect of dexamethasone, at least, on TNF mRNA. Therefore, combined/simultaneous inhibition of dexamethasone-induced multiple anti-inflammatory genes may produce a higher repressive effect. This was subsequently tested in chapter four where both DUSP1 and ZFP36 were knocked-down in combination. However, because combined knock-down of both ZFP36 and

DUSP1 did not significantly alter the percentage repression of TNF by dexamethasone, a role for additional non-DUSP1-/non-ZFP36-dependent repressive mechanisms for the repression of TNF by dexamethasone is predicated. Thus, these data show that there are likely to be many more glucocorticoid-induced effector genes, which can exert repressive effects in addition to those described here. Consequently, the repression of inflammatory genes by glucocorticoids may not be affected by a simple knock-down of one, or only a few, glucocorticoid-inducible genes.

Therefore, the data presented in chapters three and four point to the possibility that glucocorticoid- induced genes redundantly repress inflammatory gene expression, and could be a reason as to why

DUSP1- and ZFP36-specific siRNAs did not significantly reverse the dexamethasone-induced repression of TNF.

170

One of the major limitations of this work is that these findings are currently restricted to only a few specific inflammatory genes in A549 cells. Therefore, further testing with additional inflammatory genes, in different cell types, using a different inducing agent is required. However, confirmation of some of the findings in chapter four and five, with respect to the effects of IL1B and dexamethasone on DUSP1, ZFP36, TNF and IRF1 expression in primary HBE cells provides considerable support that the regulatory mechanisms described herein for the A549 cells are physiologically and therapeutically relevant. In addition, confirmation of inducibility of a number of genes, including DUSP1, RGS2, TSC22D3, FKBP51 and CDKN1C, with potential anti- inflammatory/bronchoprotective effects, by dexamethasone and budesonide in a variety of cell types, including primary HBE, primary airway smooth muscle, primary human bronchial fibroblast and primary human umbilical vein endothelial cells also supports the concept that multiple effector genes are induced upon treatment with glucocorticoids (Leigh et al. unpublished data) (280, 282, 372, 469). In addition, our laboratory has also demonstrated that the expression of TSC22D3 and FKBP51 was enhanced in asthmatics taking ICS (372). Further support for this work comes from the recent in vivo data showing that the expression of multiple anti- inflammatory/bronchoprotective genes, including DUSP1 and ZFP36, was increased in the human airways following ICS administration (Leigh et al. unpublished data). However, further characterization and functional analyses of such anti-inflammatory genes is warranted to understand the anti-inflammatory action of glucocorticoids in more detail.

171

6.4 A possible role for transrepression in the dexamethasone-induced repression of inflammatory genes

The work presented in this thesis has mainly focused on evaluating the role for glucocorticoid- inducible gene expression or transactivation on dexamethasone-induced inflammatory gene repression. However, the previous data from our laboratory indicated that inflammatory genes were subject to repression via both NR3C1 transrepression and transactivation (273). In addition, since the repression of IRF1 and TNF by dexamethasone was insensitive to the translational blocker, cycloheximide, the role for transrepression cannot be ignored (273). In this context, the expression of IL1B-induced TNF and IRF1 is NF-B dependent (chapter 5 and (273)). Thus, the obvious explanation for the observed repressive effect of dexamethasone would be a classical transrepression mechanism involving binding of NR3C1 to inflammatory transcription factors, such as NF-B. This hypothesis of transrepression in respect of TNF is further supported by the data in HeLa cells showing a significant binding of NR3C1 and NF-B to TNF promoter upon glucocorticoid and TNF co-treatment (470). Furthermore, the recent data in BEAS-2B cells suggest that TNF-induced binding of p65 (RELA) to IRF1 and TNF was modestly attenuated by dexamethasone (Dr. Anthony Gerber – personal communication), and further point to the possibility that the observed repression of IRF1 and TNF by dexamethasone may indeed involve classical NR3C1 transrepression. Whilst this may hold true, roles for miRNAs or other RNA- mediated events in the presence of dexamethasone, while completely speculative, cannot be excluded due to the fact that treatment with cycloheximide would not prevent the induction of such

RNAs by dexamethasone (273).

172

Furthermore, IL1B-induced transcription of both, TNF (chapter four) and IRF1 (chapter five), was significantly attenuated following dexamethasone treatment. This could be attributed to the effect of dexamethasone on the recruitment of RNA Pol II to target promoters or due to effects on the initiation or elongation of transcription, rate-limiting steps for gene activation (136). In this regard, the transcriptional activity of RNA Pol II is influenced by the three transcription elongation factors, namely, DSIF, NELF and P-TEFb (141). Association of DSIF and NELF with RNA Pol II triggers transcriptional pausing, whereas P-TEFb allows RNA Pol II to enter into the productive elongation phase (142, 471, 472). Furthermore, P-TEFb-mediated phosphorylation of the CTD of RNA Pol

II at S2 facilitates the release of DSIF and NELF from RNA Pol II, thereby reversing the elongation block, leading to elongation of transcription (472, 473). In the presence of glucocorticoids, this P-

TEFb-mediated phosphorylation of the Pol II CTD at S2 is inhibited, resulting in stalling of RNA

Pol II on the DNA template, which in turn may inhibit transcription (474). Thus, in the presence of glucocorticoids, even though RNA Pol II is present at the promoter, RNA pol II may lack the ability to undergo elongation of transcription and consequently resulting in inhibition of transcription by glucocorticoids. In this context, in the case of IL8, NR3C1 competed with P-TEFb for binding to NF-B and inhibited the phosphorylation of S2, thereby producing transcriptional repression (475). Thus, dexamethasone-induced transcriptional repression of IRF1 and TNF could involve mechanisms related to Pol II recruitment or inhibition of initiation or elongation of transcription. However, this needs to be further tested. Furthermore, this hypothesis about the effect of dexamethasone on RNA pol II recruitment to TNF and IRF1 promoter is further supported by the fact that in BEAS-2B cells dexamethasone modestly reduced TNF-induced RNA pol II

173

recruitment to IRF1 and TNF promoter (Dr. Anthony Gerber – personal communication). Hence, in addition to the effect on NF-B, inhibition of basic transcriptional machinery by dexamethasone could also play an important role in the repression of TNF and IRF1.

6.5 Implication for glucocorticoid therapy and new drug design

Even though ICS remains the cornerstone treatment for chronic inflammatory diseases, such as asthma (1), poor patient responses to ICS therapies, by asthmatics who smoke, or have viral infection, necessitates high, often oral, doses of glucocorticoids over prolonged periods (476). In such cases, therapeutic usefulness is compromised by systemic contraindications (477). Thus, there remains an urgent unmet clinical need to facilitate the rational design of improved anti- inflammatory NR3C1 ligands and combinatorial add-on therapies with reduced side-effect profile

(392).

Since transactivation was believed to be associated predominantly with the side-effects of glucocorticoids, pharmaceutical companies and a number of researchers sought to separate transactivation from transrepression and designed so-called dissociated steroids. These glucocorticoid ligands have the ability to transrepress, but would lack the ability to transactivate

(478). However, the data presented in this thesis clearly argue against this concept. Truly separating transactivation from transrepression may in fact reduce the anti-inflammatory potential of NR3C1 ligands, at least at early time points, by inhibiting the up-regulation of anti- inflammatory genes, such as DUSP1, by glucocorticoids. Potential dissociated glucocorticoid ligands were screened for transrepression using NF-B- and AP-1-dependent reporter systems or

174

by analysing the repression of only few selected inflammatory genes (72, 478, 479). Thus, a dissociated steroidal compound, RU24858, even though being a poor transactivator, shows anti- inflammatory properties by efficiently repressing NF-B- and AP-1-dependent transcriptional responses (478, 479). However, despite retaining their anti-inflammatory properties in vivo and in vitro, dissociated glucocorticoids were not devoid of some of the side-effects associated with conventional glucocorticoids (480). For example, RU24858 was not completely dissociated in vivo and still showed bone metabolism similar to the conventional glucocorticoid, budesonide (480). In addition, the transactivation action of most of the dissociated glucocorticoids was analysed in cell- based assays using simple GRE-based reporters (73). This will only allow the screening of ligands that cannot up-regulate the genes induced by simple GREs. In reality, NR3C1-dependent transactivation can also be induced by multiple mechanisms other than the activation of classical

GRE-dependent transcription (254). Thus, NR3C1 can interact, cross-talk or synergise with

STAT3, C/EBP and other nuclear hormone receptors to induce transcriptional activation (254, 481,

482). Additionally, the data also suggest that interaction of NR3C1 with NF-B and AP-1 also results in transcriptional activation (483, 484). In this regard, computational analysis of the promoter regions of nearly 500 genes revealed that these regions contain multiple transcription factor binding sites, including those for AP-1 and C/EBP, and further suggests that majority of

GRE sites are in fact composite elements that are bound by other transcription factors in addition to NR3C1 (485). Furthermore, mutations in either of the two NR3C1 activation domains, AF1 and

AF2, or dimerization deficient mutant (dim) of NR3C1 (GRdim) did not produce complete loss of dexamethasone-inducible genes (486). This suggests that induction of different dexamethasone-

175

inducible genes is regulated by different mechanisms of transactivation, presumably mechanisms other than at a classical GRE (254). Thus, despite the fact that GRdim mutant mice were defective in dexamethasone-induced GRE-dependent transcription (487), a number of glucocorticoid- inducible genes were still induced in a GRdim mutant mice (486). These data clearly suggest that the failure to induce simple GRE-dependent transcription may not reflect an inability to induce the expression of ‘real genes’ by NR3C1 ligands. This argument is also equally supported by the data from our laboratory showing that even though RU24858 is a dissociated glucocorticoid, RU24858 still induces the expression of anti-inflammatory DUSP1 and TSC22D3, and further questions the effect of dissociated NR3C1 ligands (271). In addition, like dexamethasone, RU24858-mediated repression of IL8 and PTGS2 steady-state mRNA and protein is also sensitive to transcriptional and translational blockade (271). These data therefore collectively suggest that glucocorticoid- inducible gene expression or transactivation plays an important role in mediating the anti- inflammatory effects of glucocorticoids. In this context, increased expression of DUSP1 by glucocorticoids inhibits MAPK activation and inflammatory gene expression, thereby is often regarded as a key anti-inflammatory effector mechanism (296, 392). Going with this notion, work presented in chapter three also shows that dexamethasone-induced DUSP1 plays a transient, but partial, role in the repression of MAPKs and inflammatory genes by dexamethasone.

One of the major contraindications associated with long-term usage of high dose glucocorticoids is severe side-effects. Thus, the holy grail of asthma therapy is to reduce the dose of glucocorticoids, so that the side-effects associated with long term glucocorticoid use can be prevented or minimised (GINA, 2009). In this regard, in the presence of LABA, a similar or higher

176

clinical efficacy of glucocorticoids can be achieved with relatively lower dosage, thus LABAs are said to be steroid sparing (82). Due to this reason, it was expected that addition of LABA may allow glucocorticoids to be used at relatively lower doses, and will have added benefits to reduce the risk of side-effects (75). Therefore, current asthma therapy recommends using the combination of LABA and ICS (GINA, 2009). In this respect, a number of glucocorticoid-induced genes, including DUSP1, CDKN1C, and RGS2, with potential anti-inflammatory/bronchoprotective effects, show enhanced expression in the presence of LABA (282). This further suggests the importance of glucocorticoid-induced gene expression. These data therefore collectively support the concept that addition of LABA to ICS therapy would likely improve the anti-inflammatory action of glucocorticoids (see chapter one for more detail). In addition, like LABAs, PDE4 inhibitors also act by increasing the intracellular levels of cAMP (488). Therefore, PDE4 inhibitors are also likely to increase DUSP1 expression (489). For example, a PDE4-selective inhibitor, rolipram, was found to enhance DUSP1 expression, and enhanced DUSP1 expression, at least partly, contributed towards the anti-inflammatory effects of rolipram (490). This observation suggests that PDE4 inhibitors may also enhance the anti-inflammatory effects of glucocorticoids.

However, current use of PDE4 inhibitors is limited due to their low therapeutic index (491).

Therefore, if PDE4 inhibitors produce beneficial effects on glucocorticoid action at a dose lower than the one that produces side-effects, they could be used in combination with ICS. Yet it remains to be seen whether this combination has any beneficial effects in asthma. Thus, identifying such treatment combinations where the therapeutic effectiveness of glucocorticoids can be enhanced

177

without using LABA may have added benefits especially in conditions where LABA cannot be utilised (492).

Despite the fact that DUSP1 is important for the inhibition of inflammatory gene expression, glucocorticoid-induced DUSP1 may also play a role in the inhibition of osteoblast proliferation and may thereby further contribute to the development of osteoporosis, a known side-effect associated with high-dose glucocorticoid treatment (477, 493). In addition, data presented in chapter five suggested that enhanced DUSP1 expression maintains IRF1 and IRF1-dependent

CXCL10 expression. Since increased expression of IRF1 has been widely associated with asthma pathogenesis and reduced glucocorticoid sensitivity (210, 407), the current findings may potentially help to explain why glucocorticoids are ineffective in viral-induced asthma exacerbations. Furthermore, such observations also indicate a long-awaited need to focus research attention to develop selective GR/NR3C1 agonists (SEGRAs) that do not enhance or up-regulate

DUSP1 and produce improved repressive effect (at least on IRF1 and IRF1-dependent responses).

Furthermore, SEGRAs may also prove to be beneficial for functional analyses of other glucocorticoid-induced genes that are negatively regulated by DUSP1 expression. In this regard, the data presented in chapter four suggested that the loss of dexamethasone-induced DUSP1 results in increased expression of ZFP36, an ARE-binding protein with potential anti-inflammatory effects. Thus, functional analysis of dexamethasone-induced ZFP36 could possibly be carried out using SEGRAs that do not up-regulate DUSP1. However, at this point, it remains to be seen whether this could ever be achieved.

178

6.6 Overall conclusion

Taken together, the data presented in this thesis clearly indicates that IL1B-induced DUSP1 and

ZFP36 play an important role in feedback and feed-forward inhibition of MAPKs and ARE- containing inflammatory transcripts, such as TNF, respectively. In addition, the data also suggest that dexamethasone-induced DUSP1 plays a transient, but partial, role in the repression of MAPKs and certain inflammatory genes by dexamethasone. Thus, the role of other glucocorticoid-induced effector mechanisms must be considered and needs to be investigated in order to guide the development of new NR3C1 ligands. In addition, although DUSP1 is essential for early onset of repression of MAPK by dexamethasone, maintenance of IRF1, through DUSP1-mediated inhibition of MAPKs, may, in part, exert insensitivity to glucocorticoids.

In summary, documenting the roles for specific glucocorticoid-inducible genes in the repression of inflammatory genes will substantially impact the rational development of novel NR3C1 modulators and combination therapies for the treatment of asthma. Moreover, the understanding of glucocorticoid-dependent mechanism(s) that may maintain the aspects of the inflammatory response could help to identify new ways of combating glucocorticoid insensitivity in severe asthma.

179

References

1. Barnes, P. J. (1998) Anti-inflammatory actions of glucocorticoids: molecular mechanisms. Clin. Sci. (Lond) 94 (6), 557-572 2. Bateman, E. D. Hurd, S. S. Barnes, P. J. Bousquet, J. Drazen, J. M. FitzGerald, M. Gibson, P. Ohta, K. O'Byrne, P. Pedersen, S. E. Pizzichini, E. Sullivan, S. D. Wenzel, S. E. Zar, H. J. (2008) Global strategy for asthma management and prevention: GINA executive summary. Eur. Respir. J. 31 (1), 143-178 3. Locksley, R. M. (2010) Asthma and allergic inflammation. Cell 140 (6), 777-783 4. To, T. Stanojevic, S. Moores, G. Gershon, A. S. Bateman, E. D. Cruz, A. A. Boulet, L. P. (2012) Global asthma prevalence in adults: findings from the cross-sectional world health survey. BMC. Public Health 12, 204 5. Vercelli, D. (2008) Discovering susceptibility genes for asthma and allergy. Nat. Rev. Immunol. 8 (3), 169-182 6. Mao, Y. Semenciw, R. Morrison, H. MacWilliam, L. Davies, J. Wigle, D. (1987) Increased rates of illness and death from asthma in Canada. CMAJ. 137 (7), 620-624 7. Senthilselvan, A. (1998) Prevalence of physician-diagnosed asthma in Saskatchewan, 1981 to 1990. Chest 114 (2), 388-392 8. Braman, S. S. (2006) The global burden of asthma. Chest 130 (1 Suppl), 4S-12S 9. Barnes, P. J. (2008) Immunology of asthma and chronic obstructive pulmonary disease. Nat. Rev. Immunol. 8 (3), 183-192 10. Busse, W. W. Lemanske, R. F., Jr. (2001) Asthma. N. Engl. J. Med. 344 (5), 350-362 11. Shifren, A. Witt, C. Christie, C. Castro, M. (2012) Mechanisms of remodeling in asthmatic airways. J. Allergy (Cairo. ) 2012, 316049 12. Burrows, B. Martinez, F. D. Halonen, M. Barbee, R. A. Cline, M. G. (1989) Association of asthma with serum IgE levels and skin-test reactivity to allergens. N. Engl. J. Med. 320 (5), 271-277 13. Hammad, H. Lambrecht, B. N. (2008) Dendritic cells and epithelial cells: linking innate and adaptive immunity in asthma. Nat. Rev. Immunol. 8 (3), 193-204 14. Jahnsen, F. L. Strickland, D. H. Thomas, J. A. Tobagus, I. T. Napoli, S. Zosky, G. R. Turner, D. J. Sly, P. D. Stumbles, P. A. Holt, P. G. (2006) Accelerated antigen sampling and transport by airway mucosal dendritic cells following inhalation of a bacterial stimulus. J. Immunol. 177 (9), 5861-5867 15. Wan, H. Winton, H. L. Soeller, C. Tovey, E. R. Gruenert, D. C. Thompson, P. J. Stewart, G. A. Taylor, G. W. Garrod, D. R. Cannell, M. B. Robinson, C. (1999) Der p 1 facilitates transepithelial allergen delivery by disruption of tight junctions. J. Clin. Invest 104 (1), 123-133 16. Bellou, A. Finn, P. W. (2005) Costimulation: critical pathways in the immunologic regulation of asthma. Curr. Allergy Asthma Rep. 5 (2), 149-154 17. Georas, S. N. Guo, J. De, F. U. Casolaro, V. (2005) T-helper cell type-2 regulation in allergic disease. Eur. Respir. J. 26 (6), 1119-1137

180

18. Gould, H. J. Sutton, B. J. (2008) IgE in allergy and asthma today. Nat. Rev. Immunol. 8 (3), 205-217 19. Beaty, S. R. Rose, C. E., Jr. Sung, S. S. (2007) Diverse and potent chemokine production by lung CD11bhigh dendritic cells in homeostasis and in allergic lung inflammation. J. Immunol. 178 (3), 1882-1895 20. Wardlaw, A. J. (1999) Molecular basis for selective eosinophil trafficking in asthma: A multistep paradigm. J. Allergy Clin. Immunol. 104 (5), 917-926 21. Sampson, A. P. (2000) The role of eosinophils and neutrophils in inflammation. Clin. Exp. Allergy 30 Suppl 1, 22-27 22. Rothenberg, M. E. (1998) Eosinophilia. N. Engl. J. Med. 338 (22), 1592-1600 23. Ordonez, C. L. Shaughnessy, T. E. Matthay, M. A. Fahy, J. V. (2000) Increased neutrophil numbers and IL-8 levels in airway secretions in acute severe asthma: Clinical and biologic significance. Am. J. Respir. Crit Care Med. 161 (4 Pt 1), 1185-1190 24. Wark, P. A. Johnston, S. L. Moric, I. Simpson, J. L. Hensley, M. J. Gibson, P. G. (2002) Neutrophil degranulation and cell lysis is associated with clinical severity in virus-induced asthma. Eur. Respir. J. 19 (1), 68-75 25. Coyle, A. J. Tsuyuki, S. Bertrand, C. Huang, S. Aguet, M. Alkan, S. S. Anderson, G. P. (1996) Mice lacking the IFN-gamma receptor have impaired ability to resolve a lung eosinophilic inflammatory response associated with a prolonged capacity of T cells to exhibit a Th2 cytokine profile. J. Immunol. 156 (8), 2680-2685 26. Gavett, S. H. O'Hearn, D. J. Li, X. Huang, S. K. Finkelman, F. D. Wills-Karp, M. (1995) Interleukin 12 inhibits antigen-induced airway hyperresponsiveness, inflammation, and Th2 cytokine expression in mice. J. Exp. Med. 182 (5), 1527-1536 27. Finn, P. W. Bigby, T. D. (2009) Innate immunity and asthma. Proc. Am. Thorac. Soc. 6 (3), 260-265 28. Corrigan, C. J. Kay, A. B. (1990) CD4 T-lymphocyte activation in acute severe asthma. Relationship to disease severity and atopic status. Am. Rev. Respir. Dis. 141 (4 Pt 1), 970- 977 29. Cembrzynska-Nowak, M. Szklarz, E. Inglot, A. D. Teodorczyk-Injeyan, J. A. (1993) Elevated release of tumor necrosis factor-alpha and interferon-gamma by bronchoalveolar leukocytes from patients with bronchial asthma. Am. Rev. Respir. Dis. 147 (2), 291-295 30. Hartnell, A. Robinson, D. S. Kay, A. B. Wardlaw, A. J. (1993) CD69 is expressed by human eosinophils activated in vivo in asthma and in vitro by cytokines. Immunology 80 (2), 281-286 31. Vock, C. Hauber, H. P. Wegmann, M. (2010) The other T helper cells in asthma pathogenesis. J. Allergy (Cairo. ) 2010, 519298 32. Jeffery, P. K. (1998) Structural and inflammatory changes in COPD: a comparison with asthma. Thorax 53 (2), 129-136 33. Baena-Cagnani, C. Rossi, G. A. Canonica, G. W. (2007) Airway remodelling in children: when does it start? Curr. Opin. Allergy Clin. Immunol. 7 (2), 196-200

181

34. Payne, D. N. Rogers, A. V. Adelroth, E. Bandi, V. Guntupalli, K. K. Bush, A. Jeffery, P. K. (2003) Early thickening of the reticular basement membrane in children with difficult asthma. Am. J. Respir. Crit Care Med. 167 (1), 78-82 35. Singanayagam, A. Joshi, P. V. Mallia, P. Johnston, S. L. (2012) Viruses exacerbating chronic pulmonary disease: the role of immune modulation. BMC. Med. 10, 27 36. Lane, S. Molina, J. Plusa, T. (2006) An international observational prospective study to determine the cost of asthma exacerbations (COAX). Respir. Med. 100 (3), 434-450 37. Spencer, S. Calverley, P. M. Sherwood, B. P. Jones, P. W. (2001) Health status deterioration in patients with chronic obstructive pulmonary disease. Am. J. Respir. Crit Care Med. 163 (1), 122-128 38. Traves, S. L. Proud, D. (2007) Viral-associated exacerbations of asthma and COPD. Curr. Opin. Pharmacol. 7 (3), 252-258 39. Lambert, H. P. Stern, H. (1972) Infective factors in exacerbations of bronchitis and asthma. Br. Med. J. 3 (5822), 323-327 40. Minor, T. E. Dick, E. C. DeMeo, A. N. Ouellette, J. J. Cohen, M. Reed, C. E. (1974) Viruses as precipitants of asthmatic attacks in children. JAMA 227 (3), 292-298 41. Johnston, S. L. Pattemore, P. K. Sanderson, G. Smith, S. Lampe, F. Josephs, L. Symington, P. O'Toole, S. Myint, S. H. Tyrrell, D. A. . (1995) Community study of role of viral infections in exacerbations of asthma in 9-11 year old children. BMJ 310 (6989), 1225- 1229 42. Nicholson, K. G. Kent, J. Ireland, D. C. (1993) Respiratory viruses and exacerbations of asthma in adults. BMJ 307 (6910), 982-986 43. Yamaya, M. (2012) Virus infection-induced bronchial asthma exacerbation. Pulm. Med. 2012, 834826 44. Papadopoulos, N. G. Papi, A. Meyer, J. Stanciu, L. A. Salvi, S. Holgate, S. T. Johnston, S. L. (2001) Rhinovirus infection up-regulates eotaxin and eotaxin-2 expression in bronchial epithelial cells. Clin. Exp. Allergy 31 (7), 1060-1066 45. Schroth, M. K. Grimm, E. Frindt, P. Galagan, D. M. Konno, S. I. Love, R. Gern, J. E. (1999) Rhinovirus replication causes RANTES production in primary bronchial epithelial cells. Am. J. Respir. Cell Mol. Biol. 20 (6), 1220-1228 46. Terajima, M. Yamaya, M. Sekizawa, K. Okinaga, S. Suzuki, T. Yamada, N. Nakayama, K. Ohrui, T. Oshima, T. Numazaki, Y. Sasaki, H. (1997) Rhinovirus infection of primary cultures of human tracheal epithelium: role of ICAM-1 and IL-1beta. Am. J. Physiol 273 (4 Pt 1), L749-L759 47. Zhu, Z. Tang, W. Gwaltney, J. M., Jr. Wu, Y. Elias, J. A. (1997) Rhinovirus stimulation of interleukin-8 in vivo and in vitro: role of NF-kappaB. Am. J. Physiol 273 (4 Pt 1), L814- L824 48. Jackson, D. J. Johnston, S. L. (2010) The role of viruses in acute exacerbations of asthma. J. Allergy Clin. Immunol. 125 (6), 1178-1187 49. Contoli, M. Message, S. D. Laza-Stanca, V. Edwards, M. R. Wark, P. A. Bartlett, N. W. Kebadze, T. Mallia, P. Stanciu, L. A. Parker, H. L. Slater, L. Lewis-Antes, A. Kon, O. M. Holgate, S. T. Davies, D. E. Kotenko, S. V. Papi, A. Johnston, S. L. (2006) Role of

182

deficient type III interferon-lambda production in asthma exacerbations. Nat. Med. 12 (9), 1023-1026 50. Wark, P. A. Johnston, S. L. Bucchieri, F. Powell, R. Puddicombe, S. Laza-Stanca, V. Holgate, S. T. Davies, D. E. (2005) Asthmatic bronchial epithelial cells have a deficient innate immune response to infection with rhinovirus. J. Exp. Med. 201 (6), 937-947 51. Laza-Stanca, V. Stanciu, L. A. Message, S. D. Edwards, M. R. Gern, J. E. Johnston, S. L. (2006) Rhinovirus replication in human macrophages induces NF-kappaB-dependent tumor necrosis factor alpha production. J. Virol. 80 (16), 8248-8258 52. Bochkov, Y. A. Hanson, K. M. Keles, S. Brockman-Schneider, R. A. Jarjour, N. N. Gern, J. E. (2010) Rhinovirus-induced modulation of gene expression in bronchial epithelial cells from subjects with asthma. Mucosal. Immunol. 3 (1), 69-80 53. Lopez-Souza, N. Favoreto, S. Wong, H. Ward, T. Yagi, S. Schnurr, D. Finkbeiner, W. E. Dolganov, G. M. Widdicombe, J. H. Boushey, H. A. Avila, P. C. (2009) In vitro susceptibility to rhinovirus infection is greater for bronchial than for nasal airway epithelial cells in human subjects. J. Allergy Clin. Immunol. 123 (6), 1384-1390 54. Thomson, N. C. Chaudhuri, R. Livingston, E. (2004) Asthma and cigarette smoking. Eur. Respir. J. 24 (5), 822-833 55. Venarske, D. L. Busse, W. W. Griffin, M. R. Gebretsadik, T. Shintani, A. K. Minton, P. A. Peebles, R. S. Hamilton, R. Weisshaar, E. Vrtis, R. Higgins, S. B. Hartert, T. V. (2006) The relationship of rhinovirus-associated asthma hospitalizations with inhaled corticosteroids and smoking. J. Infect. Dis. 193 (11), 1536-1543 56. Crystal, R. G. Randell, S. H. Engelhardt, J. F. Voynow, J. Sunday, M. E. (2008) Airway epithelial cells: current concepts and challenges. Proc. Am. Thorac. Soc. 5 (7), 772-777 57. Knight, D. A. Holgate, S. T. (2003) The airway epithelium: structural and functional properties in health and disease. Respirology. 8 (4), 432-446 58. Thompson, A. B. Robbins, R. A. Romberger, D. J. Sisson, J. H. Spurzem, J. R. Teschler, H. Rennard, S. I. (1995) Immunological functions of the pulmonary epithelium. Eur. Respir. J. 8 (1), 127-149 59. Williams, M. C. (2003) Alveolar type I cells: molecular phenotype and development. Annu. Rev. Physiol 65, 669-695 60. Mason, R. J. (2006) Biology of alveolar type II cells. Respirology. 11 Suppl, S12-S15 61. Wang, Y. Bai, C. Li, K. Adler, K. B. Wang, X. (2008) Role of airway epithelial cells in development of asthma and allergic rhinitis. Respir. Med. 102 (7), 949-955 62. Barnes, P. J. (2001) Molecular mechanisms of corticosteroids in allergic diseases. Allergy 56 (10), 928-936 63. Ying, S. O'Connor, B. Ratoff, J. Meng, Q. Mallett, K. Cousins, D. Robinson, D. Zhang, G. Zhao, J. Lee, T. H. Corrigan, C. (2005) Thymic stromal lymphopoietin expression is increased in asthmatic airways and correlates with expression of Th2-attracting chemokines and disease severity. J. Immunol. 174 (12), 8183-8190 64. Zhou, B. Comeau, M. R. De, S. T. Liggitt, H. D. Dahl, M. E. Lewis, D. B. Gyarmati, D. Aye, T. Campbell, D. J. Ziegler, S. F. (2005) Thymic stromal lymphopoietin as a key initiator of allergic airway inflammation in mice. Nat. Immunol. 6 (10), 1047-1053

183

65. Reber, L. Da Silva, C. A. Frossard, N. (2006) Stem cell factor and its receptor c-Kit as targets for inflammatory diseases. Eur. J. Pharmacol. 533 (1-3), 327-340 66. Diamond, G. Legarda, D. Ryan, L. K. (2000) The innate immune response of the respiratory epithelium. Immunol. Rev. 173, 27-38 67. Schleimer, R. P. (2004) Glucocorticoids suppress inflammation but spare innate immune responses in airway epithelium. Proc. Am. Thorac. Soc. 1 (3), 222-230 68. Holgate, S. T. Polosa, R. (2008) Treatment strategies for allergy and asthma. Nat. Rev. Immunol. 8 (3), 218-230 69. Barnes, P. J. Adcock, I. M. (1995) Steroid resistance in asthma. QJM. 88 (7), 455-468 70. Szefler, S. J. Leung, D. Y. (1997) Glucocorticoid-resistant asthma: pathogenesis and clinical implications for management. Eur. Respir. J. 10 (7), 1640-1647 71. Barnes, P. J. (2004) New drugs for asthma. Nat. Rev. Drug Discov. 3 (10), 831-844 72. Newton, R. (2000) Molecular mechanisms of glucocorticoid action: what is important? Thorax 55 (7), 603-613 73. Uings, I. J. Farrow, S. N. (2005) A pharmacological approach to enhancing the therapeutic index of corticosteroids in airway inflammatory disease. Curr. Opin. Pharmacol. 5 (3), 221-226 74. Guilbert, T. W. Morgan, W. J. Zeiger, R. S. Mauger, D. T. Boehmer, S. J. Szefler, S. J. Bacharier, L. B. Lemanske, R. F., Jr. Strunk, R. C. Allen, D. B. Bloomberg, G. R. Heldt, G. Krawiec, M. Larsen, G. Liu, A. H. Chinchilli, V. M. Sorkness, C. A. Taussig, L. M. Martinez, F. D. (2006) Long-term inhaled corticosteroids in preschool children at high risk for asthma. N. Engl. J. Med. 354 (19), 1985-1997 75. Giembycz, M. A. Kaur, M. Leigh, R. Newton, R. (2008) A Holy Grail of asthma management: toward understanding how long-acting beta(2)-adrenoceptor agonists enhance the clinical efficacy of inhaled corticosteroids. Br. J. Pharmacol. 153 (6), 1090- 1104 76. Johnson, M. (1998) The beta-adrenoceptor. Am. J. Respir. Crit Care Med. 158 (5 Pt 3), S146-S153 77. Johnson, M. (2004) Interactions between corticosteroids and beta2-agonists in asthma and chronic obstructive pulmonary disease. Proc. Am. Thorac. Soc. 1 (3), 200-206 78. Giembycz, M. A. Newton, R. (2006) Beyond the dogma: novel beta2-adrenoceptor signalling in the airways. Eur. Respir. J. 27 (6), 1286-1306 79. Barnes, P. J. (2007) Scientific rationale for using a single inhaler for asthma control. Eur. Respir. J. 29 (3), 587-595 80. Bousquet, J. (2000) Global initiative for asthma (GINA) and its objectives. Clin. Exp. Allergy 30 Suppl 1, 2-5 81. Miller-Larsson, A. Selroos, O. (2006) Advances in asthma and COPD treatment: combination therapy with inhaled corticosteroids and long-acting beta 2-agonists. Curr. Pharm. Des 12 (25), 3261-3279 82. Greening, A. P. Ind, P. W. Northfield, M. Shaw, G. (1994) Added salmeterol versus higher-dose corticosteroid in asthma patients with symptoms on existing inhaled corticosteroid. Allen & Hanburys Limited UK Study Group. Lancet 344 (8917), 219-224

184

83. Mak, J. C. Nishikawa, M. Barnes, P. J. (1995) Glucocorticosteroids increase beta 2- adrenergic receptor transcription in human lung. Am. J. Physiol 268 (1 Pt 1), L41-L46 84. Chong, L. K. Drury, D. E. Dummer, J. F. Ghahramani, P. Schleimer, R. P. Peachell, P. T. (1997) Protection by dexamethasone of the functional desensitization to beta 2- adrenoceptor-mediated responses in human lung mast cells. Br. J. Pharmacol. 121 (4), 717-722 85. Kalavantavanich, K. Schramm, C. M. (2000) Dexamethasone potentiates high-affinity beta-agonist binding and g(s)alpha protein expression in airway smooth muscle. Am. J. Physiol Lung Cell Mol. Physiol 278 (5), L1101-L1106 86. Moore, P. E. Laporte, J. D. Gonzalez, S. Moller, W. Heyder, J. Panettieri, R. A., Jr. Shore, S. A. (1999) Glucocorticoids ablate IL-1beta-induced beta-adrenergic hyporesponsiveness in human airway smooth muscle cells. Am. J. Physiol 277 (5 Pt 1), L932-L942 87. Pang, L. Holland, E. Knox, A. J. (1998) Role of cyclo-oxygenase-2 induction in interleukin-1beta induced attenuation of cultured human airway smooth muscle cell cyclic AMP generation in response to isoprenaline. Br. J. Pharmacol. 125 (6), 1320-1328 88. Pauwels, R. A. Lofdahl, C. G. Postma, D. S. Tattersfield, A. E. O'Byrne, P. Barnes, P. J. Ullman, A. (1997) Effect of inhaled formoterol and budesonide on exacerbations of asthma. Formoterol and Corticosteroids Establishing Therapy (FACET) International Study Group. N. Engl. J. Med. 337 (20), 1405-1411 89. Nie, M. Knox, A. J. Pang, L. (2005) beta2-Adrenoceptor agonists, like glucocorticoids, repress eotaxin gene transcription by selective inhibition of histone H4 acetylation. J. Immunol. 175 (1), 478-486 90. Barnes, P. J. (2011) Biochemical basis of asthma therapy. J. Biol. Chem. 286 (38), 32899- 32905 91. Hallstrand, T. S. Henderson, W. R., Jr. (2010) An update on the role of leukotrienes in asthma. Curr. Opin. Allergy Clin. Immunol. 10 (1), 60-66 92. Guan, K. L. (1994) The mitogen activated protein kinase signal transduction pathway: from the cell surface to the nucleus. Cell Signal. 6 (6), 581-589 93. Chen, Z. Gibson, T. B. Robinson, F. Silvestro, L. Pearson, G. Xu, B. Wright, A. Vanderbilt, C. Cobb, M. H. (2001) MAP kinases. Chem. Rev. 101 (8), 2449-2476 94. Minden, A. Karin, M. (1997) Regulation and function of the JNK subgroup of MAP kinases. Biochim. Biophys. Acta 1333 (2), F85-104 95. Raman, M. Chen, W. Cobb, M. H. (2007) Differential regulation and properties of MAPKs. Oncogene 26 (22), 3100-3112 96. Minden, A. Karin, M. (1997) Regulation and function of the JNK subgroup of MAP kinases. Biochim. Biophys. Acta 1333 (2), F85-104 97. Johnson, G. L. Lapadat, R. (2002) Mitogen-activated protein kinase pathways mediated by ERK, JNK, and p38 protein kinases. Science 298 (5600), 1911-1912 98. Cargnello, M. Roux, P. P. (2011) Activation and function of the MAPKs and their substrates, the MAPK-activated protein kinases. Microbiol. Mol. Biol. Rev. 75 (1), 50-83 99. Krishna, M. Narang, H. (2008) The complexity of mitogen-activated protein kinases (MAPKs) made simple. Cell Mol. Life Sci. 65 (22), 3525-3544

185

100. Boutros, T. Chevet, E. Metrakos, P. (2008) Mitogen-activated protein (MAP) kinase/MAP kinase phosphatase regulation: roles in cell growth, death, and cancer. Pharmacol. Rev. 60 (3), 261-310 101. Pearson, G. Robinson, F. Beers, G. T. Xu, B. E. Karandikar, M. Berman, K. Cobb, M. H. (2001) Mitogen-activated protein (MAP) kinase pathways: regulation and physiological functions. Endocr. Rev. 22 (2), 153-183 102. Roux, P. P. Blenis, J. (2004) ERK and p38 MAPK-activated protein kinases: a family of protein kinases with diverse biological functions. Microbiol. Mol. Biol. Rev. 68 (2), 320- 344 103. Gao, M. Labuda, T. Xia, Y. Gallagher, E. Fang, D. Liu, Y. C. Karin, M. (2004) Jun turnover is controlled through JNK-dependent phosphorylation of the E3 ligase Itch. Science 306 (5694), 271-275 104. Keyse, S. M. (1998) Protein phosphatases and the regulation of MAP kinase activity. Semin. Cell Dev. Biol. 9 (2), 143-152 105. Keyse, S. M. (2000) Protein phosphatases and the regulation of mitogen-activated protein kinase signalling. Curr. Opin. Cell Biol. 12 (2), 186-192 106. Alessi, D. R. Gomez, N. Moorhead, G. Lewis, T. Keyse, S. M. Cohen, P. (1995) Inactivation of p42 MAP kinase by protein phosphatase 2A and a protein tyrosine phosphatase, but not CL100, in various cell lines. Curr. Biol. 5 (3), 283-295 107. Pulido, R. Zuniga, A. Ullrich, A. (1998) PTP-SL and STEP protein tyrosine phosphatases regulate the activation of the extracellular signal-regulated kinases ERK1 and ERK2 by association through a kinase interaction motif. EMBO J. 17 (24), 7337-7350 108. Guan, K. L. Broyles, S. S. Dixon, J. E. (1991) A Tyr/Ser protein phosphatase encoded by vaccinia virus. Nature 350 (6316), 359-362 109. Patterson, K. I. Brummer, T. O'Brien, P. M. Daly, R. J. (2009) Dual-specificity phosphatases: critical regulators with diverse cellular targets. Biochem. J. 418 (3), 475-489 110. Schulze-Osthoff, K. Ferrari, D. Riehemann, K. Wesselborg, S. (1997) Regulation of NF- kappa B activation by MAP kinase cascades. Immunobiology 198 (1-3), 35-49 111. Farooq, A. Zhou, M. M. (2004) Structure and regulation of MAPK phosphatases. Cell Signal. 16 (7), 769-779 112. Hu, J. H. Chen, T. Zhuang, Z. H. Kong, L. Yu, M. C. Liu, Y. Zang, J. W. Ge, B. X. (2007) Feedback control of MKP-1 expression by p38. Cell Signal. 19 (2), 393-400 113. Duan, W. Wong, W. S. (2006) Targeting mitogen-activated protein kinases for asthma. Curr. Drug Targets. 7 (6), 691-698 114. King, E. M. Holden, N. S. Gong, W. Rider, C. F. Newton, R. (2009) Inhibition of NF- kappaB-dependent transcription by MKP-1: transcriptional repression by glucocorticoids occurring via p38 MAPK. J. Biol. Chem. 284 (39), 26803-26815 115. Karin, M. (2006) Nuclear factor-kappaB in cancer development and progression. Nature 441 (7092), 431-436 116. Buxade, M. Parra-Palau, J. L. Proud, C. G. (2008) The Mnks: MAP kinase-interacting kinases (MAP kinase signal-integrating kinases). Front Biosci. 13, 5359-5373

186

117. Ronkina, N. Kotlyarov, A. Gaestel, M. (2008) MK2 and MK3--a pair of isoenzymes? Front Biosci. 13, 5511-5521 118. Kim, C. Sano, Y. Todorova, K. Carlson, B. A. Arpa, L. Celada, A. Lawrence, T. Otsu, K. Brissette, J. L. Arthur, J. S. Park, J. M. (2008) The kinase p38 alpha serves cell type- specific inflammatory functions in skin injury and coordinates pro- and anti-inflammatory gene expression. Nat. Immunol. 9 (9), 1019-1027 119. Puig-Kroger, A. Relloso, M. Fernandez-Capetillo, O. Zubiaga, A. Silva, A. Bernabeu, C. Corbi, A. L. (2001) Extracellular signal-regulated protein kinase signaling pathway negatively regulates the phenotypic and functional maturation of monocyte-derived human dendritic cells. Blood 98 (7), 2175-2182 120. Boehme, S. A. Sullivan, S. K. Crowe, P. D. Santos, M. Conlon, P. J. Sriramarao, P. Bacon, K. B. (1999) Activation of mitogen-activated protein kinase regulates eotaxin-induced eosinophil migration. J. Immunol. 163 (3), 1611-1618 121. Ishizuka, T. Okajima, F. Ishiwara, M. Iizuka, K. Ichimonji, I. Kawata, T. Tsukagoshi, H. Dobashi, K. Nakazawa, T. Mori, M. (2001) Sensitized mast cells migrate toward the antigen: a response regulated by p38 mitogen-activated protein kinase and Rho-associated coiled-coil-forming protein kinase. J. Immunol. 167 (4), 2298-2304 122. Murphy, L. O. Smith, S. Chen, R. H. Fingar, D. C. Blenis, J. (2002) Molecular interpretation of ERK signal duration by immediate early gene products. Nat. Cell Biol. 4 (8), 556-564 123. Sabapathy, K. Hochedlinger, K. Nam, S. Y. Bauer, A. Karin, M. Wagner, E. F. (2004) Distinct roles for JNK1 and JNK2 in regulating JNK activity and c-Jun-dependent cell proliferation. Mol. Cell 15 (5), 713-725 124. Barnes, P. J. Adcock, I. M. (1998) Transcription factors and asthma. Eur. Respir. J. 12 (1), 221-234 125. Wuyts, W. A. Vanaudenaerde, B. M. Dupont, L. J. Demedts, M. G. Verleden, G. M. (2003) Involvement of p38 MAPK, JNK, p42/p44 ERK and NF-kappaB in IL-1beta-induced chemokine release in human airway smooth muscle cells. Respir. Med. 97 (7), 811-817 126. Yang, Z. Zhang, X. Darrah, P. A. Mosser, D. M. (2010) The regulation of Th1 responses by the p38 MAPK. J. Immunol. 185 (10), 6205-6213 127. Zhu, J. Wu, X. Goel, S. Gowda, N. M. Kumar, S. Krishnegowda, G. Mishra, G. Weinberg, R. Li, G. Gaestel, M. Muta, T. Gowda, D. C. (2009) MAPK-activated protein kinase 2 differentially regulates plasmodium falciparum glycosylphosphatidylinositol-induced production of tumor necrosis factor-{alpha} and interleukin-12 in macrophages. J. Biol. Chem. 284 (23), 15750-15761 128. Korhonen, R. Huotari, N. Hommo, T. Leppanen, T. Moilanen, E. (2012) The expression of interleukin-12 is increased by MAP kinase phosphatase-1 through a mechanism related to interferon regulatory factor 1. Mol. Immunol. 129. Zaheer, R. S. Koetzler, R. Holden, N. S. Wiehler, S. Proud, D. (2009) Selective transcriptional down-regulation of human rhinovirus-induced production of CXCL10 from airway epithelial cells via the MEK1 pathway. J. Immunol. 182 (8), 4854-4864

187

130. Medoff, B. D. Sauty, A. Tager, A. M. Maclean, J. A. Smith, R. N. Mathew, A. Dufour, J. H. Luster, A. D. (2002) IFN-gamma-inducible protein 10 (CXCL10) contributes to airway hyperreactivity and airway inflammation in a mouse model of asthma. J. Immunol. 168 (10), 5278-5286 131. Orphanides, G. Reinberg, D. (2002) A unified theory of gene expression. Cell 108 (4), 439-451 132. Luger, K. Mader, A. W. Richmond, R. K. Sargent, D. F. Richmond, T. J. (1997) Crystal structure of the nucleosome core particle at 2.8 A resolution. Nature 389 (6648), 251-260 133. Kadonaga, J. T. (2004) Regulation of RNA polymerase II transcription by sequence- specific DNA binding factors. Cell 116 (2), 247-257 134. Strahl, B. D. Allis, C. D. (2000) The language of covalent histone modifications. Nature 403 (6765), 41-45 135. Saha, A. Wittmeyer, J. Cairns, B. R. (2006) Chromatin remodelling: the industrial revolution of DNA around histones. Nat. Rev. Mol. Cell Biol. 7 (6), 437-447 136. Medzhitov, R. Horng, T. (2009) Transcriptional control of the inflammatory response. Nat. Rev. Immunol. 9 (10), 692-703 137. Svejstrup, J. Q. (2004) The RNA polymerase II transcription cycle: cycling through chromatin. Biochim. Biophys. Acta 1677 (1-3), 64-73 138. Baumann, M. Pontiller, J. Ernst, W. (2010) Structure and basal transcription complex of RNA polymerase II core promoters in the mammalian genome: an overview. Mol. Biotechnol. 45 (3), 241-247 139. Smale, S. T. Kadonaga, J. T. (2003) The RNA polymerase II core promoter. Annu. Rev. Biochem. 72, 449-479 140. Hahn, S. (2004) Structure and mechanism of the RNA polymerase II transcription machinery. Nat. Struct. Mol. Biol. 11 (5), 394-403 141. Uptain, S. M. Kane, C. M. Chamberlin, M. J. (1997) Basic mechanisms of transcript elongation and its regulation. Annu. Rev. Biochem. 66, 117-172 142. Wada, T. Takagi, T. Yamaguchi, Y. Watanabe, D. Handa, H. (1998) Evidence that P- TEFb alleviates the negative effect of DSIF on RNA polymerase II-dependent transcription in vitro. EMBO J. 17 (24), 7395-7403 143. Saunders, A. Core, L. J. Lis, J. T. (2006) Breaking barriers to transcription elongation. Nat. Rev. Mol. Cell Biol. 7 (8), 557-567 144. Proudfoot, N. J. Furger, A. Dye, M. J. (2002) Integrating mRNA processing with transcription. Cell 108 (4), 501-512 145. Barnes, P. J. (2006) Transcription factors in airway diseases. Lab Invest 86 (9), 867-872 146. Sen, R. Baltimore, D. (1986) Multiple nuclear factors interact with the immunoglobulin enhancer sequences. Cell 46 (5), 705-716 147. Grilli, M. Chiu, J. J. Lenardo, M. J. (1993) NF-kappa B and Rel: participants in a multiform transcriptional regulatory system. Int. Rev. Cytol. 143, 1-62 148. Grimm, S. Baeuerle, P. A. (1993) The inducible transcription factor NF-kappa B: structure-function relationship of its protein subunits. Biochem. J. 290 ( Pt 2), 297-308

188

149. Gilmore, T. D. (1997) Introduction: The Rel/NF-kappaB signal transduction pathway. Semin. Cancer Biol. 8 (2), 61-62 150. Schmitz, M. L. Mattioli, I. Buss, H. Kracht, M. (2004) NF-kappaB: a multifaceted transcription factor regulated at several levels. Chembiochem. 5 (10), 1348-1358 151. Karin, M. Ben-Neriah, Y. (2000) Phosphorylation meets ubiquitination: the control of NF- [kappa]B activity. Annu. Rev. Immunol. 18, 621-663 152. Pahl, H. L. (1999) Activators and target genes of Rel/NF-kappaB transcription factors. Oncogene 18 (49), 6853-6866 153. Karin, M. Ben-Neriah, Y. (2000) Phosphorylation meets ubiquitination: the control of NF- [kappa]B activity. Annu. Rev. Immunol. 18, 621-663 154. Luo, J. L. Kamata, H. Karin, M. (2005) IKK/NF-kappaB signaling: balancing life and death--a new approach to cancer therapy. J. Clin. Invest 115 (10), 2625-2632 155. Hayden, M. S. Ghosh, S. (2004) Signaling to NF-kappaB. Genes Dev. 18 (18), 2195-2224 156. Pereira, S. G. Oakley, F. (2008) Nuclear factor-kappaB1: regulation and function. Int. J. Biochem. Cell Biol. 40 (8), 1425-1430 157. Gilmore, T. D. (2006) Introduction to NF-kappaB: players, pathways, perspectives. Oncogene 25 (51), 6680-6684 158. Cheng, Q. Cant, C. A. Moll, T. Hofer-Warbinek, R. Wagner, E. Birnstiel, M. L. Bach, F. H. de, M. R. (1994) NK-kappa B subunit-specific regulation of the I kappa B alpha promoter. J. Biol. Chem. 269 (18), 13551-13557 159. Ito, C. Y. Kazantsev, A. G. Baldwin, A. S., Jr. (1994) Three NF-kappa B sites in the I kappa B-alpha promoter are required for induction of gene expression by TNF alpha. Nucleic Acids Res. 22 (18), 3787-3792 160. Chen, L. Fischle, W. Verdin, E. Greene, W. C. (2001) Duration of nuclear NF-kappaB action regulated by reversible acetylation. Science 293 (5535), 1653-1657 161. Klement, J. F. Rice, N. R. Car, B. D. Abbondanzo, S. J. Powers, G. D. Bhatt, P. H. Chen, C. H. Rosen, C. A. Stewart, C. L. (1996) IkappaBalpha deficiency results in a sustained NF-kappaB response and severe widespread dermatitis in mice. Mol. Cell Biol. 16 (5), 2341-2349 162. Rahman, I. MacNee, W. (1998) Role of transcription factors in inflammatory lung diseases. Thorax 53 (7), 601-612 163. Caramori, G. Oates, T. Nicholson, A. G. Casolari, P. Ito, K. Barnes, P. J. Papi, A. Adcock, I. M. Chung, K. F. (2009) Activation of NF-kappaB transcription factor in asthma death. Histopathology 54 (4), 507-509 164. Hart, L. A. Krishnan, V. L. Adcock, I. M. Barnes, P. J. Chung, K. F. (1998) Activation and localization of transcription factor, nuclear factor-kappaB, in asthma. Am. J. Respir. Crit Care Med. 158 (5 Pt 1), 1585-1592 165. Vignola, A. M. Chiappara, G. Siena, L. Bruno, A. Gagliardo, R. Merendino, A. M. Polla, B. S. Arrigo, A. P. Bonsignore, G. Bousquet, J. Chanez, P. (2001) Proliferation and activation of bronchial epithelial cells in corticosteroid-dependent asthma. J. Allergy Clin. Immunol. 108 (5), 738-746

189

166. Gagliardo, R. Chanez, P. Mathieu, M. Bruno, A. Costanzo, G. Gougat, C. Vachier, I. Bousquet, J. Bonsignore, G. Vignola, A. M. (2003) Persistent activation of nuclear factor- kappaB signaling pathway in severe uncontrolled asthma. Am. J. Respir. Crit Care Med. 168 (10), 1190-1198 167. Funkhouser, A. W. Kang, J. A. Tan, A. Li, J. Zhou, L. Abe, M. K. Solway, J. Hershenson, M. B. (2004) Rhinovirus 16 3C protease induces interleukin-8 and granulocyte- macrophage colony-stimulating factor expression in human bronchial epithelial cells. Pediatr. Res. 55 (1), 13-18 168. Papi, A. Johnston, S. L. (1999) Respiratory epithelial cell expression of vascular cell adhesion molecule-1 and its up-regulation by rhinovirus infection via NF-kappaB and GATA transcription factors. J. Biol. Chem. 274 (42), 30041-30051 169. Pestka, S. Langer, J. A. Zoon, K. C. Samuel, C. E. (1987) Interferons and their actions. Annu. Rev. Biochem. 56, 727-777 170. Taniguchi, T. Ogasawara, K. Takaoka, A. Tanaka, N. (2001) IRF family of transcription factors as regulators of host defense. Annu. Rev. Immunol. 19, 623-655 171. Nguyen, H. Hiscott, J. Pitha, P. M. (1997) The growing family of interferon regulatory factors. Cytokine Growth Factor Rev. 8 (4), 293-312 172. Tamura, T. Yanai, H. Savitsky, D. Taniguchi, T. (2008) The IRF family transcription factors in immunity and oncogenesis. Annu. Rev. Immunol. 26, 535-584 173. Uegaki, K. Shirakawa, M. Harada, H. Taniguchi, T. Kyogoku, Y. (1995) Secondary structure and folding topology of the DNA binding domain of interferon regulatory factor 2, as revealed by NMR spectroscopy. FEBS Lett. 359 (2-3), 184-188 174. Fujii, Y. Shimizu, T. Kusumoto, M. Kyogoku, Y. Taniguchi, T. Hakoshima, T. (1999) Crystal structure of an IRF-DNA complex reveals novel DNA recognition and cooperative binding to a tandem repeat of core sequences. EMBO J. 18 (18), 5028-5041 175. Chen, W. Royer, W. E., Jr. (2010) Structural insights into interferon regulatory factor activation. Cell Signal. 22 (6), 883-887 176. Ozato, K. Tailor, P. Kubota, T. (2007) The interferon regulatory factor family in host defense: mechanism of action. J. Biol. Chem. 282 (28), 20065-20069 177. Fujita, T. Kimura, Y. Miyamoto, M. Barsoumian, E. L. Taniguchi, T. (1989) Induction of endogenous IFN-alpha and IFN-beta genes by a regulatory transcription factor, IRF-1. Nature 337 (6204), 270-272 178. Abdollahi, A. Lord, K. A. Hoffman-Liebermann, B. Liebermann, D. A. (1991) Interferon regulatory factor 1 is a myeloid differentiation primary response gene induced by interleukin 6 and leukemia inhibitory factor: role in growth inhibition. Cell Growth Differ. 2 (8), 401-407 179. Fujita, T. Reis, L. F. Watanabe, N. Kimura, Y. Taniguchi, T. Vilcek, J. (1989) Induction of the transcription factor IRF-1 and interferon-beta mRNAs by cytokines and activators of second-messenger pathways. Proc. Natl. Acad. Sci. U. S. A 86 (24), 9936-9940 180. Harada, H. Fujita, T. Miyamoto, M. Kimura, Y. Maruyama, M. Furia, A. Miyata, T. Taniguchi, T. (1989) Structurally similar but functionally distinct factors, IRF-1 and IRF-

190

2, bind to the same regulatory elements of IFN and IFN-inducible genes. Cell 58 (4), 729- 739 181. Harada, H. Takahashi, E. Itoh, S. Harada, K. Hori, T. A. Taniguchi, T. (1994) Structure and regulation of the human interferon regulatory factor 1 (IRF-1) and IRF-2 genes: implications for a gene network in the interferon system. Mol. Cell Biol. 14 (2), 1500-1509 182. Miyamoto, M. Fujita, T. Kimura, Y. Maruyama, M. Harada, H. Sudo, Y. Miyata, T. Taniguchi, T. (1988) Regulated expression of a gene encoding a nuclear factor, IRF-1, that specifically binds to IFN-beta gene regulatory elements. Cell 54 (6), 903-913 183. Li, X. Leung, S. Qureshi, S. Darnell, J. E., Jr. Stark, G. R. (1996) Formation of STAT1- STAT2 heterodimers and their role in the activation of IRF-1 gene transcription by interferon-alpha. J. Biol. Chem. 271 (10), 5790-5794 184. Kroger, A. Koster, M. Schroeder, K. Hauser, H. Mueller, P. P. (2002) Activities of IRF-1. J. Interferon Cytokine Res. 22 (1), 5-14 185. Dou, L. Liang, H. F. Geller, D. A. Chen, Y. F. Chen, X. P. (2014) The regulation role of interferon regulatory factor-1 gene and clinical relevance. Hum. Immunol. 75 (11), 1110- 1114 186. Sanceau, J. Hiscott, J. Delattre, O. Wietzerbin, J. (2000) IFN-beta induces serine phosphorylation of Stat-1 in Ewing's sarcoma cells and mediates apoptosis via induction of IRF-1 and activation of caspase-7. Oncogene 19 (30), 3372-3383 187. Dror, N. ter-Koltunoff, M. Azriel, A. Amariglio, N. Jacob-Hirsch, J. Zeligson, S. Morgenstern, A. Tamura, T. Hauser, H. Rechavi, G. Ozato, K. Levi, B. Z. (2007) Identification of IRF-8 and IRF-1 target genes in activated macrophages. Mol. Immunol. 44 (4), 338-346 188. Zaheer, R. S. Proud, D. (2010) Human rhinovirus-induced epithelial production of CXCL10 is dependent upon IFN regulatory factor-1. Am. J. Respir. Cell Mol. Biol. 43 (4), 413-421 189. Farrar, M. A. Schreiber, R. D. (1993) The molecular cell biology of interferon-gamma and its receptor. Annu. Rev. Immunol. 11, 571-611 190. Mosmann, T. R. Coffman, R. L. (1989) TH1 and TH2 cells: different patterns of lymphokine secretion lead to different functional properties. Annu. Rev. Immunol. 7, 145- 173 191. McElligott, D. L. Phillips, J. A. Stillman, C. A. Koch, R. J. Mosier, D. E. Hobbs, M. V. (1997) CD4+ T cells from IRF-1-deficient mice exhibit altered patterns of cytokine expression and cell subset homeostasis. J. Immunol. 159 (9), 4180-4186 192. Taki, S. Sato, T. Ogasawara, K. Fukuda, T. Sato, M. Hida, S. Suzuki, G. Mitsuyama, M. Shin, E. H. Kojima, S. Taniguchi, T. Asano, Y. (1997) Multistage regulation of Th1-type immune responses by the transcription factor IRF-1. Immunity. 6 (6), 673-679 193. Elser, B. Lohoff, M. Kock, S. Giaisi, M. Kirchhoff, S. Krammer, P. H. Li-Weber, M. (2002) IFN-gamma represses IL-4 expression via IRF-1 and IRF-2. Immunity. 17 (6), 703- 712

191

194. Maruyama, S. Kanoh, M. Matsumoto, A. Kuwahara, M. Yamashita, M. Asano, Y. (2015) A novel function of interferon regulatory factor-1: inhibition of Th2 cells by down- regulating the Il4 gene during Listeria infection. Int. Immunol. 27 (3), 143-152 195. Colonna, M. (2007) TLR pathways and IFN-regulatory factors: to each its own. Eur. J. Immunol. 37 (2), 306-309 196. Kamijo, R. Harada, H. Matsuyama, T. Bosland, M. Gerecitano, J. Shapiro, D. Le, J. Koh, S. I. Kimura, T. Green, S. J. . (1994) Requirement for transcription factor IRF-1 in NO synthase induction in macrophages. Science 263 (5153), 1612-1615 197. Martin, E. Nathan, C. Xie, Q. W. (1994) Role of interferon regulatory factor 1 in induction of nitric oxide synthase. J. Exp. Med. 180 (3), 977-984 198. Paludan, S. R. Malmgaard, L. Ellermann-Eriksen, S. Bosca, L. Mogensen, S. C. (2001) Interferon (IFN)-gamma and Herpes simplex virus/tumor necrosis factor-alpha synergistically induce nitric oxide synthase 2 in macrophages through cooperative action of nuclear factor-kappa B and IFN regulatory factor-1. Eur. Cytokine Netw. 12 (2), 297- 308 199. Koetzler, R. Zaheer, R. S. Wiehler, S. Holden, N. S. Giembycz, M. A. Proud, D. (2009) Nitric oxide inhibits human rhinovirus-induced transcriptional activation of CXCL10 in airway epithelial cells. J. Allergy Clin. Immunol. 123 (1), 201-208 200. Sanders, S. P. Siekierski, E. S. Porter, J. D. Richards, S. M. Proud, D. (1998) Nitric oxide inhibits rhinovirus-induced cytokine production and viral replication in a human respiratory epithelial cell line. J. Virol. 72 (2), 934-942 201. Sanders, S. P. Kim, J. Connolly, K. R. Porter, J. D. Siekierski, E. S. Proud, D. (2001) Nitric oxide inhibits rhinovirus-induced granulocyte macrophage colony-stimulating factor production in bronchial epithelial cells. Am. J. Respir. Cell Mol. Biol. 24 (3), 317-325 202. MacMicking, J. Xie, Q. W. Nathan, C. (1997) Nitric oxide and macrophage function. Annu. Rev. Immunol. 15, 323-350 203. Lohoff, M. Ferrick, D. Mittrucker, H. W. Duncan, G. S. Bischof, S. Rollinghoff, M. Mak, T. W. (1997) Interferon regulatory factor-1 is required for a T helper 1 immune response in vivo. Immunity. 6 (6), 681-689 204. Wang, T. N. Chu, Y. T. Chen, W. Y. Feng, W. W. Shih, N. H. Hsiang, C. H. Ko, Y. C. (2006) Association of interferon-gamma and interferon regulatory factor 1 polymorphisms with asthma in a family-based association study in Taiwan. Clin. Exp. Allergy 36 (9), 1147- 1152 205. Schedel, M. Pinto, L. A. Schaub, B. Rosenstiel, P. Cherkasov, D. Cameron, L. Klopp, N. Illig, T. Vogelberg, C. Weiland, S. K. von, M. E. Lohoff, M. Kabesch, M. (2008) IRF-1 gene variations influence IgE regulation and atopy. Am. J. Respir. Crit Care Med. 177 (6), 613-621 206. Nakao, F. Ihara, K. Kusuhara, K. Sasaki, Y. Kinukawa, N. Takabayashi, A. Nishima, S. Hara, T. (2001) Association of IFN-gamma and IFN regulatory factor 1 polymorphisms with childhood atopic asthma. J. Allergy Clin. Immunol. 107 (3), 499-504

192

207. Noguchi, E. Shibasaki, M. Arinami, T. Yamakawa-Kobayashi, K. Yokouchi, Y. Takeda, K. Matsui, A. Hamaguchi, H. (2000) Mutation screening of interferon regulatory factor 1 gene (IRF-1) as a candidate gene for atopy/asthma. Clin. Exp. Allergy 30 (11), 1562-1567 208. Walley, A. J. Wiltshire, S. Ellis, C. M. Cookson, W. O. (2001) Linkage and allelic association of 5 cytokine cluster genetic markers with atopy and asthma associated traits. Genomics 72 (1), 15-20 209. Bhandare, R. Damera, G. Banerjee, A. Flammer, J. R. Keslacy, S. Rogatsky, I. Panettieri, R. A. Amrani, Y. Tliba, O. (2010) Glucocorticoid receptor interacting protein-1 restores glucocorticoid responsiveness in steroid-resistant airway structural cells. Am. J. Respir. Cell Mol. Biol. 42 (1), 9-15 210. Tliba, O. Damera, G. Banerjee, A. Gu, S. Baidouri, H. Keslacy, S. Amrani, Y. (2008) Cytokines induce an early steroid resistance in airway smooth muscle cells: novel role of interferon regulatory factor-1. Am. J. Respir. Cell Mol. Biol. 38 (4), 463-472 211. Hamilton, T. Novotny, M. Pavicic, P. J., Jr. Herjan, T. Hartupee, J. Sun, D. Zhao, C. Datta, S. (2010) Diversity in post-transcriptional control of neutrophil chemoattractant cytokine gene expression. Cytokine 52 (1-2), 116-122 212. Wilusz, C. J. Wilusz, J. (2004) Bringing the role of mRNA decay in the control of gene expression into focus. Trends Genet. 20 (10), 491-497 213. Gingerich, T. J. Feige, J. J. LaMarre, J. (2004) AU-rich elements and the control of gene expression through regulated mRNA stability. Anim Health Res. Rev. 5 (1), 49-63 214. Anderson, P. (2008) Post-transcriptional control of cytokine production. Nat. Immunol. 9 (4), 353-359 215. Barreau, C. Paillard, L. Osborne, H. B. (2005) AU-rich elements and associated factors: are there unifying principles? Nucleic Acids Res. 33 (22), 7138-7150 216. Khabar, K. S. (2010) Post-transcriptional control during chronic inflammation and cancer: a focus on AU-rich elements. Cell Mol. Life Sci. 67 (17), 2937-2955 217. Palanisamy, V. Park, N. J. Wang, J. Wong, D. T. (2008) AUF1 and HuR proteins stabilize interleukin-8 mRNA in human saliva. J. Dent. Res. 87 (8), 772-776 218. Palanisamy, V. Jakymiw, A. Van Tubergen, E. A. D'Silva, N. J. Kirkwood, K. L. (2012) Control of cytokine mRNA expression by RNA-binding proteins and microRNAs. J. Dent. Res. 91 (7), 651-658 219. Carballo, E. Gilkeson, G. S. Blackshear, P. J. (1997) Bone marrow transplantation reproduces the tristetraprolin-deficiency syndrome in recombination activating gene-2 (-/- ) mice. Evidence that monocyte/macrophage progenitors may be responsible for TNFalpha overproduction. J. Clin. Invest 100 (5), 986-995 220. Dean, J. L. Wait, R. Mahtani, K. R. Sully, G. Clark, A. R. Saklatvala, J. (2001) The 3' untranslated region of tumor necrosis factor alpha mRNA is a target of the mRNA- stabilizing factor HuR. Mol. Cell Biol. 21 (3), 721-730 221. Malter, J. S. (1989) Identification of an AUUUA-specific messenger RNA binding protein. Science 246 (4930), 664-666 222. Taylor, G. A. Carballo, E. Lee, D. M. Lai, W. S. Thompson, M. J. Patel, D. D. Schenkman, D. I. Gilkeson, G. S. Broxmeyer, H. E. Haynes, B. F. Blackshear, P. J. (1996) A

193

pathogenetic role for TNF alpha in the syndrome of cachexia, arthritis, and autoimmunity resulting from tristetraprolin (TTP) deficiency. Immunity. 4 (5), 445-454 223. Winzen, R. Thakur, B. K. ttrich-Breiholz, O. Shah, M. Redich, N. Dhamija, S. Kracht, M. Holtmann, H. (2007) Functional analysis of KSRP interaction with the AU-rich element of interleukin-8 and identification of inflammatory mRNA targets. Mol. Cell Biol. 27 (23), 8388-8400 224. Garneau, N. L. Wilusz, J. Wilusz, C. J. (2007) The highways and byways of mRNA decay. Nat. Rev. Mol. Cell Biol. 8 (2), 113-126 225. Clement, S. L. Scheckel, C. Stoecklin, G. Lykke-Andersen, J. (2011) Phosphorylation of tristetraprolin by MK2 impairs AU-rich element mRNA decay by preventing deadenylase recruitment. Mol. Cell Biol. 31 (2), 256-266 226. Sandler, H. Stoecklin, G. (2008) Control of mRNA decay by phosphorylation of tristetraprolin. Biochem. Soc. Trans. 36 (Pt 3), 491-496 227. Winzen, R. Kracht, M. Ritter, B. Wilhelm, A. Chen, C. Y. Shyu, A. B. Muller, M. Gaestel, M. Resch, K. Holtmann, H. (1999) The p38 MAP kinase pathway signals for cytokine- induced mRNA stabilization via MAP kinase-activated protein kinase 2 and an AU-rich region-targeted mechanism. EMBO J. 18 (18), 4969-4980 228. Wilson, G. M. Lu, J. Sutphen, K. Sun, Y. Huynh, Y. Brewer, G. (2003) Regulation of A + U-rich element-directed mRNA turnover involving reversible phosphorylation of AUF1. J. Biol. Chem. 278 (35), 33029-33038 229. Zhang, W. Wagner, B. J. Ehrenman, K. Schaefer, A. W. DeMaria, C. T. Crater, D. DeHaven, K. Long, L. Brewer, G. (1993) Purification, characterization, and cDNA cloning of an AU-rich element RNA-binding protein, AUF1. Mol. Cell Biol. 13 (12), 7652-7665 230. Tsai, M. J. O'Malley, B. W. (1994) Molecular mechanisms of action of steroid/thyroid receptor superfamily members. Annu. Rev. Biochem. 63, 451-486 231. Fardella, C. E. Miller, W. L. (1996) Molecular biology of mineralocorticoid metabolism. Annu. Rev. Nutr. 16, 443-470 232. Funder, J. W. Sheppard, K. (1987) Adrenocortical steroids and the brain. Annu. Rev. Physiol 49, 397-411 233. Miller, W. L. (1988) Molecular biology of steroid hormone synthesis. Endocr. Rev. 9 (3), 295-318 234. Reichardt, H. M. Schutz, G. (1998) Glucocorticoid signalling--multiple variations of a common theme. Mol. Cell Endocrinol. 146 (1-2), 1-6 235. Balsalobre, A. Brown, S. A. Marcacci, L. Tronche, F. Kellendonk, C. Reichardt, H. M. Schutz, G. Schibler, U. (2000) Resetting of circadian time in peripheral tissues by glucocorticoid signaling. Science 289 (5488), 2344-2347 236. Jacobson, L. (2005) Hypothalamic-pituitary-adrenocortical axis regulation. Endocrinol. Metab Clin. North Am. 34 (2), 271-92, vii 237. Moras, D. Gronemeyer, H. (1998) The nuclear receptor ligand-binding domain: structure and function. Curr. Opin. Cell Biol. 10 (3), 384-391

194

238. McKay, L. I. Cidlowski, J. A. (1999) Molecular control of immune/inflammatory responses: interactions between nuclear factor-kappa B and steroid receptor-signaling pathways. Endocr. Rev. 20 (4), 435-459 239. Hollenberg, S. M. Weinberger, C. Ong, E. S. Cerelli, G. Oro, A. Lebo, R. Thompson, E. B. Rosenfeld, M. G. Evans, R. M. (1985) Primary structure and expression of a functional human glucocorticoid receptor cDNA. Nature 318 (6047), 635-641 240. Bamberger, C. M. Bamberger, A. M. de, C. M. Chrousos, G. P. (1995) Glucocorticoid receptor beta, a potential endogenous inhibitor of glucocorticoid action in humans. J. Clin. Invest 95 (6), 2435-2441 241. Oakley, R. H. Jewell, C. M. Yudt, M. R. Bofetiado, D. M. Cidlowski, J. A. (1999) The dominant negative activity of the human glucocorticoid receptor beta isoform. Specificity and mechanisms of action. J. Biol. Chem. 274 (39), 27857-27866 242. Giguere, V. Hollenberg, S. M. Rosenfeld, M. G. Evans, R. M. (1986) Functional domains of the human glucocorticoid receptor. Cell 46 (5), 645-652 243. Nicolaides, N. C. Galata, Z. Kino, T. Chrousos, G. P. Charmandari, E. (2010) The human glucocorticoid receptor: molecular basis of biologic function. Steroids 75 (1), 1-12 244. Pratt, W. B. Toft, D. O. (1997) Steroid receptor interactions with heat shock protein and immunophilin chaperones. Endocr. Rev. 18 (3), 306-360 245. Smith, D. F. Whitesell, L. Katsanis, E. (1998) Molecular chaperones: biology and prospects for pharmacological intervention. Pharmacol. Rev. 50 (4), 493-514 246. Cato, A. C. Mink, S. (2001) BAG-1 family of cochaperones in the modulation of nuclear receptor action. J. Steroid Biochem. Mol. Biol. 78 (5), 379-388 247. Barnes, P. J. Adcock, I. M. (2003) How do corticosteroids work in asthma? Ann. Intern. Med. 139 (5 Pt 1), 359-370 248. Davies, T. H. Ning, Y. M. Sanchez, E. R. (2002) A new first step in activation of steroid receptors: hormone-induced switching of FKBP51 and FKBP52 immunophilins. J. Biol. Chem. 277 (7), 4597-4600 249. Beato, M. Chavez, S. Truss, M. (1996) Transcriptional regulation by steroid hormones. Steroids 61 (4), 240-251 250. Schaaf, M. J. Cidlowski, J. A. (2002) Molecular mechanisms of glucocorticoid action and resistance. J. Steroid Biochem. Mol. Biol. 83 (1-5), 37-48 251. bbinante-Nissen, J. M. Simpson, L. G. Leikauf, G. D. (1995) Corticosteroids increase secretory leukocyte protease inhibitor transcript levels in airway epithelial cells. Am. J. Physiol 268 (4 Pt 1), L601-L606 252. Re, F. Muzio, M. De, R. M. Polentarutti, N. Giri, J. G. Mantovani, A. Colotta, F. (1994) The type II "receptor" as a decoy target for interleukin 1 in polymorphonuclear leukocytes: characterization of induction by dexamethasone and ligand binding properties of the released decoy receptor. J. Exp. Med. 179 (2), 739-743 253. Drouin, J. Sun, Y. L. Chamberland, M. Gauthier, Y. De, L. A. Nemer, M. Schmidt, T. J. (1993) Novel glucocorticoid receptor complex with DNA element of the hormone- repressed POMC gene. EMBO J. 12 (1), 145-156

195

254. Newton, R. Holden, N. S. (2007) Separating transrepression and transactivation: a distressing divorce for the glucocorticoid receptor? Mol. Pharmacol. 72 (4), 799-809 255. Surjit, M. Ganti, K. P. Mukherji, A. Ye, T. Hua, G. Metzger, D. Li, M. Chambon, P. (2011) Widespread negative response elements mediate direct repression by agonist-liganded glucocorticoid receptor. Cell 145 (2), 224-241 256. Ramamoorthy, S. Cidlowski, J. A. (2013) Ligand-induced repression of the glucocorticoid receptor gene is mediated by an NCoR1 repression complex formed by long-range chromatin interactions with intragenic glucocorticoid response elements. Mol. Cell Biol. 33 (9), 1711-1722 257. Hayashi, R. Wada, H. Ito, K. Adcock, I. M. (2004) Effects of glucocorticoids on gene transcription. Eur. J. Pharmacol. 500 (1-3), 51-62 258. Beck, I. M. Vanden, B. W. Vermeulen, L. Yamamoto, K. R. Haegeman, G. De, B. K. (2009) Crosstalk in inflammation: the interplay of glucocorticoid receptor-based mechanisms and kinases and phosphatases. Endocr. Rev. 30 (7), 830-882 259. De, B. K. Vanden, B. W. Haegeman, G. (2003) The interplay between the glucocorticoid receptor and nuclear factor-kappaB or activator protein-1: molecular mechanisms for gene repression. Endocr. Rev. 24 (4), 488-522 260. Ito, K. Barnes, P. J. Adcock, I. M. (2000) Glucocorticoid receptor recruitment of histone deacetylase 2 inhibits interleukin-1beta-induced histone H4 acetylation on lysines 8 and 12. Mol. Cell Biol. 20 (18), 6891-6903 261. Ito, K. Jazrawi, E. Cosio, B. Barnes, P. J. Adcock, I. M. (2001) p65-activated histone acetyltransferase activity is repressed by glucocorticoids: mifepristone fails to recruit HDAC2 to the p65-HAT complex. J. Biol. Chem. 276 (32), 30208-30215 262. Luecke, H. F. Yamamoto, K. R. (2005) The glucocorticoid receptor blocks P-TEFb recruitment by NFkappaB to effect promoter-specific transcriptional repression. Genes Dev. 19 (9), 1116-1127 263. Rogatsky, I. Luecke, H. F. Leitman, D. C. Yamamoto, K. R. (2002) Alternate surfaces of transcriptional coregulator GRIP1 function in different glucocorticoid receptor activation and repression contexts. Proc. Natl. Acad. Sci. U. S. A 99 (26), 16701-16706 264. Clark, A. R. Lasa, M. (2003) Crosstalk between glucocorticoids and mitogen-activated protein kinase signalling pathways. Curr. Opin. Pharmacol. 3 (4), 404-411 265. De, B. K. Haegeman, G. (2009) Minireview: latest perspectives on antiinflammatory actions of glucocorticoids. Mol. Endocrinol. 23 (3), 281-291 266. Catley, M. C. Chivers, J. E. Holden, N. S. Barnes, P. J. Newton, R. (2005) Validation of IKK beta as therapeutic target in airway inflammatory disease by adenoviral-mediated delivery of dominant-negative IKK beta to pulmonary epithelial cells. Br. J. Pharmacol. 145 (1), 114-122 267. Chan, A. Y. Vreede, F. T. Smith, M. Engelhardt, O. G. Fodor, E. (2006) Influenza virus inhibits RNA polymerase II elongation. Virology 351 (1), 210-217 268. Newton, R. Hart, L. A. Stevens, D. A. Bergmann, M. Donnelly, L. E. Adcock, I. M. Barnes, P. J. (1998) Effect of dexamethasone on interleukin-1beta-(IL-1beta)-induced nuclear

196

factor-kappaB (NF-kappaB) and kappaB-dependent transcription in epithelial cells. Eur. J. Biochem. 254 (1), 81-89 269. Chivers, J. E. Cambridge, L. M. Catley, M. C. Mak, J. C. Donnelly, L. E. Barnes, P. J. Newton, R. (2004) Differential effects of RU486 reveal distinct mechanisms for glucocorticoid repression of prostaglandin E release. Eur. J. Biochem. 271 (20), 4042-4052 270. Chang, M. M. Juarez, M. Hyde, D. M. Wu, R. (2001) Mechanism of dexamethasone- mediated interleukin-8 gene suppression in cultured airway epithelial cells. Am. J. Physiol Lung Cell Mol. Physiol 280 (1), L107-L115 271. Chivers, J. E. Gong, W. King, E. M. Seybold, J. Mak, J. C. Donnelly, L. E. Holden, N. S. Newton, R. (2006) Analysis of the dissociated steroid RU24858 does not exclude a role for inducible genes in the anti-inflammatory actions of glucocorticoids. Mol. Pharmacol. 70 (6), 2084-2095 272. Newton, R. Seybold, J. Kuitert, L. M. Bergmann, M. Barnes, P. J. (1998) Repression of cyclooxygenase-2 and prostaglandin E2 release by dexamethasone occurs by transcriptional and post-transcriptional mechanisms involving loss of polyadenylated mRNA. J. Biol. Chem. 273 (48), 32312-32321 273. King, E. M. Chivers, J. E. Rider, C. F. Minnich, A. Giembycz, M. A. Newton, R. (2013) Glucocorticoid repression of inflammatory gene expression shows differential responsiveness by transactivation- and transrepression-dependent mechanisms. PLoS. One. 8 (1), e53936 274. Wissink, S. van Heerde, E. C. vand der, B. B. van der Saag, P. T. (1998) A dual mechanism mediates repression of NF-kappaB activity by glucocorticoids. Mol. Endocrinol. 12 (3), 355-363 275. Hiragun, T. Peng, Z. Beaven, M. A. (2005) Dexamethasone up-regulates the inhibitory adaptor protein Dok-1 and suppresses downstream activation of the mitogen-activated protein kinase pathway in antigen-stimulated RBL-2H3 mast cells. Mol. Pharmacol. 67 (3), 598-603 276. Hiragun, T. Peng, Z. Beaven, M. A. (2006) Cutting edge: dexamethasone negatively regulates Syk in mast cells by up-regulating SRC-like adaptor protein. J. Immunol. 177 (4), 2047-2050 277. Nguyen, C. H. Watts, V. J. (2006) Dexamethasone-induced Ras protein 1 negatively regulates protein kinase C delta: implications for adenylyl cyclase 2 signaling. Mol. Pharmacol. 69 (5), 1763-1771 278. Chen, L. C. Zhang, Z. Myers, A. C. Huang, S. K. (2001) Cutting edge: altered pulmonary eosinophilic inflammation in mice deficient for Clara cell secretory 10-kDa protein. J. Immunol. 167 (6), 3025-3028 279. Young, J. D. Lawrence, A. J. MacLean, A. G. Leung, B. P. McInnes, I. B. Canas, B. Pappin, D. J. Stevenson, R. D. (1999) Thymosin beta 4 sulfoxide is an anti-inflammatory agent generated by monocytes in the presence of glucocorticoids. Nat. Med. 5 (12), 1424-1427 280. Holden, N. S. Bell, M. J. Rider, C. F. King, E. M. Gaunt, D. D. Leigh, R. Johnson, M. Siderovski, D. P. Heximer, S. P. Giembycz, M. A. Newton, R. (2011) beta2-Adrenoceptor

197

agonist-induced RGS2 expression is a genomic mechanism of bronchoprotection that is enhanced by glucocorticoids. Proc. Natl. Acad. Sci. U. S. A 108 (49), 19713-19718 281. Abraham, S. M. Lawrence, T. Kleiman, A. Warden, P. Medghalchi, M. Tuckermann, J. Saklatvala, J. Clark, A. R. (2006) Antiinflammatory effects of dexamethasone are partly dependent on induction of dual specificity phosphatase 1. J. Exp. Med. 203 (8), 1883-1889 282. Kaur, M. Chivers, J. E. Giembycz, M. A. Newton, R. (2008) Long-acting beta2- adrenoceptor agonists synergistically enhance glucocorticoid-dependent transcription in human airway epithelial and smooth muscle cells. Mol. Pharmacol. 73 (1), 203-214 283. Newton, R. King, E. M. Gong, W. Rider, C. F. Staples, K. J. Holden, N. S. Bergmann, M. W. (2010) Glucocorticoids inhibit IL-1beta-induced GM-CSF expression at multiple levels: roles for the ERK pathway and repression by MKP-1. Biochem. J. 427 (1), 113-124 284. Turpeinen, T. Nieminen, R. Moilanen, E. Korhonen, R. (2010) Mitogen-Activated Protein Kinase Phosphatase-1 Negatively Regulates the Expression of Interleukin-6, Interleukin- 8, and Cyclooxygenase-2 in A549 Human Lung Epithelial Cells. Journal of Pharmacology and Experimental Therapeutics 333 (1), 310-318 285. Wang, X. Liu, Y. (2007) Regulation of innate immune response by MAP kinase phosphatase-1. Cell Signal. 19 (7), 1372-1382 286. Brondello, J. M. Brunet, A. Pouyssegur, J. McKenzie, F. R. (1997) The dual specificity mitogen-activated protein kinase phosphatase-1 and -2 are induced by the p42/p44MAPK cascade. J. Biol. Chem. 272 (2), 1368-1376 287. Brondello, J. M. Pouyssegur, J. McKenzie, F. R. (1999) Reduced MAP kinase phosphatase-1 degradation after p42/p44MAPK-dependent phosphorylation. Science 286 (5449), 2514-2517 288. Chi, H. Flavell, R. A. (2008) Acetylation of MKP-1 and the control of inflammation. Sci. Signal. 1 (41), e44 289. Lin, Y. W. Chuang, S. M. Yang, J. L. (2003) ERK1/2 Achieves Sustained Activation by Stimulating MAPK Phosphatase-1 Degradation via the Ubiquitin-Proteasome Pathway. Journal of Biological Chemistry 278 (24), 21534-21541 290. Lin, Y. W. Yang, J. L. (2006) Cooperation of ERK and SCFSkp2 for MKP-1 destruction provides a positive feedback regulation of proliferating signaling. J. Biol. Chem. 281 (2), 915-926 291. Kassel, O. Sancono, A. Kratzschmar, J. Kreft, B. Stassen, M. Cato, A. C. (2001) Glucocorticoids inhibit MAP kinase via increased expression and decreased degradation of MKP-1. EMBO J. 20 (24), 7108-7116 292. Cao, W. Bao, C. Padalko, E. Lowenstein, C. J. (2008) Acetylation of mitogen-activated protein kinase phosphatase-1 inhibits Toll-like receptor signaling. J. Exp. Med. 205 (6), 1491-1503 293. Manetsch, M. Che, W. Seidel, P. Chen, Y. Ammit, A. J. (2012) MKP-1: a negative feedback effector that represses MAPK-mediated pro-inflammatory signaling pathways and cytokine secretion in human airway smooth muscle cells. Cell Signal. 24 (4), 907-913 294. Abraham, S. M. Clark, A. R. (2006) Dual-specificity phosphatase 1: a critical regulator of innate immune responses. Biochem. Soc. Trans. 34 (Pt 6), 1018-1023

198

295. Clark, A. R. (2007) Anti-inflammatory functions of glucocorticoid-induced genes. Mol. Cell Endocrinol. 275 (1-2), 79-97 296. Clark, A. R. Martins, J. R. Tchen, C. R. (2008) Role of dual specificity phosphatases in biological responses to glucocorticoids. J. Biol. Chem. 283 (38), 25765-25769 297. Lawan, A. Shi, H. Gatzke, F. Bennett, A. M. (2013) Diversity and specificity of the mitogen-activated protein kinase phosphatase-1 functions. Cell Mol. Life Sci. 70 (2), 223- 237 298. Clark, A. R. Dean, J. L. Saklatvala, J. (2003) Post-transcriptional regulation of gene expression by mitogen-activated protein kinase p38. FEBS Lett. 546 (1), 37-44 299. Ghosh, S. Hayden, M. S. (2008) New regulators of NF-kappaB in inflammation. Nat. Rev. Immunol. 8 (11), 837-848 300. Saccani, S. Pantano, S. Natoli, G. (2002) p38-Dependent marking of inflammatory genes for increased NF-kappa B recruitment. Nat. Immunol. 3 (1), 69-75 301. Wesselborg, S. Bauer, M. K. Vogt, M. Schmitz, M. L. Schulze-Osthoff, K. (1997) Activation of transcription factor NF-kappaB and p38 mitogen-activated protein kinase is mediated by distinct and separate stress effector pathways. J. Biol. Chem. 272 (19), 12422- 12429 302. Liu, Y. Gorospe, M. Yang, C. Holbrook, N. J. (1995) Role of mitogen-activated protein kinase phosphatase during the cellular response to genotoxic stress. Inhibition of c-Jun N- terminal kinase activity and AP-1-dependent gene activation. J. Biol. Chem. 270 (15), 8377-8380 303. Zhao, Q. Shepherd, E. G. Manson, M. E. Nelin, L. D. Sorokin, A. Liu, Y. (2005) The role of mitogen-activated protein kinase phosphatase-1 in the response of alveolar macrophages to lipopolysaccharide: attenuation of proinflammatory cytokine biosynthesis via feedback control of p38. J. Biol. Chem. 280 (9), 8101-8108 304. Zhao, Q. Wang, X. Nelin, L. D. Yao, Y. Matta, R. Manson, M. E. Baliga, R. S. Meng, X. Smith, C. V. Bauer, J. A. Chang, C. H. Liu, Y. (2006) MAP kinase phosphatase 1 controls innate immune responses and suppresses endotoxic shock. J. Exp. Med. 203 (1), 131-140 305. Aslam, N. Zaheer, I. (2011) The biosynthesis characteristics of TTP and TNF can be regulated through a posttranscriptional molecular loop. J. Biol. Chem. 286 (5), 3767-3776 306. Smoak, K. Cidlowski, J. A. (2006) Glucocorticoids regulate tristetraprolin synthesis and posttranscriptionally regulate tumor necrosis factor alpha inflammatory signaling. Mol. Cell Biol. 26 (23), 9126-9135 307. Jalonen, U. Leppanen, T. Kankaanranta, H. Moilanen, E. (2007) Salbutamol increases tristetraprolin expression in macrophages. Life Sci. 81 (25-26), 1651-1658 308. King, E. M. Kaur, M. Gong, W. Rider, C. F. Holden, N. S. Newton, R. (2009) Regulation of tristetraprolin expression by interleukin-1 beta and dexamethasone in human pulmonary epithelial cells: roles for nuclear factor-kappa B and p38 mitogen-activated protein kinase. J. Pharmacol. Exp. Ther. 330 (2), 575-585 309. Mahtani, K. R. Brook, M. Dean, J. L. Sully, G. Saklatvala, J. Clark, A. R. (2001) Mitogen- activated protein kinase p38 controls the expression and posttranslational modification of

199

tristetraprolin, a regulator of tumor necrosis factor alpha mRNA stability. Mol. Cell Biol. 21 (19), 6461-6469 310. Brook, M. Tchen, C. R. Santalucia, T. McIlrath, J. Arthur, J. S. Saklatvala, J. Clark, A. R. (2006) Posttranslational regulation of tristetraprolin subcellular localization and protein stability by p38 mitogen-activated protein kinase and extracellular signal-regulated kinase pathways. Mol. Cell Biol. 26 (6), 2408-2418 311. Stoecklin, G. Stubbs, T. Kedersha, N. Wax, S. Rigby, W. F. Blackwell, T. K. Anderson, P. (2004) MK2-induced tristetraprolin:14-3-3 complexes prevent stress granule association and ARE-mRNA decay. EMBO J. 23 (6), 1313-1324 312. Sun, L. Stoecklin, G. Van, W. S. Hinkovska-Galcheva, V. Guo, R. F. Anderson, P. Shanley, T. P. (2007) Tristetraprolin (TTP)-14-3-3 complex formation protects TTP from dephosphorylation by protein phosphatase 2a and stabilizes tumor necrosis factor-alpha mRNA. J. Biol. Chem. 282 (6), 3766-3777 313. Brooks, S. A. Blackshear, P. J. (2013) Tristetraprolin (TTP): interactions with mRNA and proteins, and current thoughts on mechanisms of action. Biochim. Biophys. Acta 1829 (6- 7), 666-679 314. Hau, H. H. Walsh, R. J. Ogilvie, R. L. Williams, D. A. Reilly, C. S. Bohjanen, P. R. (2007) Tristetraprolin recruits functional mRNA decay complexes to ARE sequences. J. Cell Biochem. 100 (6), 1477-1492 315. Lykke-Andersen, J. Wagner, E. (2005) Recruitment and activation of mRNA decay enzymes by two ARE-mediated decay activation domains in the proteins TTP and BRF-1. Genes Dev. 19 (3), 351-361 316. Lai, W. S. Kennington, E. A. Blackshear, P. J. (2003) Tristetraprolin and its family members can promote the cell-free deadenylation of AU-rich element-containing mRNAs by poly(A) ribonuclease. Mol. Cell Biol. 23 (11), 3798-3812 317. Ishmael, F. T. Fang, X. Galdiero, M. R. Atasoy, U. Rigby, W. F. Gorospe, M. Cheadle, C. Stellato, C. (2008) Role of the RNA-binding protein tristetraprolin in glucocorticoid- mediated gene regulation. J. Immunol. 180 (12), 8342-8353 318. Shi, J. X. Li, J. S. Hu, R. Shi, Y. Su, X. Guo, X. J. Li, X. M. (2014) Tristetraprolin is involved in the glucocorticoid-mediated interleukin 8 repression. Int. Immunopharmacol. 22 (2), 480-485 319. Jalonen, U. Lahti, A. Korhonen, R. Kankaanranta, H. Moilanen, E. (2005) Inhibition of tristetraprolin expression by dexamethasone in activated macrophages. Biochem. Pharmacol. 69 (5), 733-740 320. Huotari, N. Hommo, T. Taimi, V. Nieminen, R. Moilanen, E. Korhonen, R. (2012) Regulation of tristetraprolin expression by mitogen-activated protein kinase phosphatase- 1. APMIS 120 (12), 988-999 321. Barnes, P. J. Adcock, I. M. (2009) Glucocorticoid resistance in inflammatory diseases. Lancet 373 (9678), 1905-1917 322. Keenan, C. R. Salem, S. Fietz, E. R. Gualano, R. C. Stewart, A. G. (2012) Glucocorticoid- resistant asthma and novel anti-inflammatory drugs. Drug Discov. Today 17 (17-18), 1031- 1038

200

323. Berry, M. A. Hargadon, B. Shelley, M. Parker, D. Shaw, D. E. Green, R. H. Bradding, P. Brightling, C. E. Wardlaw, A. J. Pavord, I. D. (2006) Evidence of a role of tumor necrosis factor alpha in refractory asthma. N. Engl. J. Med. 354 (7), 697-708 324. Morjaria, J. B. Chauhan, A. J. Babu, K. S. Polosa, R. Davies, D. E. Holgate, S. T. (2008) The role of a soluble TNFalpha receptor fusion protein (etanercept) in corticosteroid refractory asthma: a double blind, randomised, placebo controlled trial. Thorax 63 (7), 584- 591 325. Howarth, P. H. Babu, K. S. Arshad, H. S. Lau, L. Buckley, M. McConnell, W. Beckett, P. Al, A. M. Chauhan, A. Wilson, S. J. Reynolds, A. Davies, D. E. Holgate, S. T. (2005) Tumour necrosis factor (TNFalpha) as a novel therapeutic target in symptomatic corticosteroid dependent asthma. Thorax 60 (12), 1012-1018 326. Franchimont, D. Martens, H. Hagelstein, M. T. Louis, E. Dewe, W. Chrousos, G. P. Belaiche, J. Geenen, V. (1999) Tumor necrosis factor alpha decreases, and interleukin-10 increases, the sensitivity of human monocytes to dexamethasone: potential regulation of the glucocorticoid receptor. J. Clin. Endocrinol. Metab 84 (8), 2834-2839 327. Sullivan, D. E. Ferris, M. Pociask, D. Brody, A. R. (2005) Tumor necrosis factor-alpha induces transforming growth factor-beta1 expression in lung fibroblasts through the extracellular signal-regulated kinase pathway. Am. J. Respir. Cell Mol. Biol. 32 (4), 342- 349 328. Robins, S. Roussel, L. Schachter, A. Risse, P. A. Mogas, A. K. Olivenstein, R. Martin, J. G. Hamid, Q. Rousseau, S. (2011) Steroid-insensitive ERK1/2 activity drives CXCL8 synthesis and neutrophilia by airway smooth muscle. Am. J. Respir. Cell Mol. Biol. 45 (5), 984-990 329. Heaton, T. Rowe, J. Turner, S. Aalberse, R. C. de, K. N. Suriyaarachchi, D. Serralha, M. Holt, B. J. Hollams, E. Yerkovich, S. Holt, K. Sly, P. D. Goldblatt, J. Le, S. P. Holt, P. G. (2005) An immunoepidemiological approach to asthma: identification of in-vitro T-cell response patterns associated with different wheezing phenotypes in children. Lancet 365 (9454), 142-149 330. Yang, M. Kumar, R. K. Foster, P. S. (2009) Pathogenesis of steroid-resistant airway hyperresponsiveness: interaction between IFN-gamma and TLR4/MyD88 pathways. J. Immunol. 182 (8), 5107-5115 331. Li, J. J. Wang, W. Baines, K. J. Bowden, N. A. Hansbro, P. M. Gibson, P. G. Kumar, R. K. Foster, P. S. Yang, M. (2010) IL-27/IFN-gamma induce MyD88-dependent steroid- resistant airway hyperresponsiveness by inhibiting glucocorticoid signaling in macrophages. J. Immunol. 185 (7), 4401-4409 332. Salem, S. Harris, T. Mok, J. S. Li, M. Y. Keenan, C. R. Schuliga, M. J. Stewart, A. G. (2012) Transforming growth factor-beta impairs glucocorticoid activity in the A549 lung adenocarcinoma cell line. Br. J. Pharmacol. 166 (7), 2036-2048 333. Ismaili, N. Garabedian, M. J. (2004) Modulation of glucocorticoid receptor function via phosphorylation. Ann. N. Y. Acad. Sci. 1024, 86-101 334. Goleva, E. Kisich, K. O. Leung, D. Y. (2002) A role for STAT5 in the pathogenesis of IL- 2-induced glucocorticoid resistance. J. Immunol. 169 (10), 5934-5940

201

335. Loke, T. K. Mallett, K. H. Ratoff, J. O'Connor, B. J. Ying, S. Meng, Q. Soh, C. Lee, T. H. Corrigan, C. J. (2006) Systemic glucocorticoid reduces bronchial mucosal activation of activator protein 1 components in glucocorticoid-sensitive but not glucocorticoid-resistant asthmatic patients. J. Allergy Clin. Immunol. 118 (2), 368-375 336. Hakonarson, H. Bjornsdottir, U. S. Halapi, E. Bradfield, J. Zink, F. Mouy, M. Helgadottir, H. Gudmundsdottir, A. S. Andrason, H. Adalsteinsdottir, A. E. Kristjansson, K. Birkisson, I. Arnason, T. Andresdottir, M. Gislason, D. Gislason, T. Gulcher, J. R. Stefansson, K. (2005) Profiling of genes expressed in peripheral blood mononuclear cells predicts glucocorticoid sensitivity in asthma patients. Proc. Natl. Acad. Sci. U. S. A 102 (41), 14789-14794 337. Barnes, P. J. (2010) Mechanisms and resistance in glucocorticoid control of inflammation. J. Steroid Biochem. Mol. Biol. 120 (2-3), 76-85 338. Ito, K. Caramori, G. Adcock, I. M. (2007) Therapeutic potential of phosphatidylinositol 3-kinase inhibitors in inflammatory respiratory disease. J. Pharmacol. Exp. Ther. 321 (1), 1-8 339. Barnes, P. J. (2009) Role of HDAC2 in the pathophysiology of COPD. Annu. Rev. Physiol 71, 451-464 340. Hew, M. Bhavsar, P. Torrego, A. Meah, S. Khorasani, N. Barnes, P. J. Adcock, I. Chung, K. F. (2006) Relative corticosteroid insensitivity of peripheral blood mononuclear cells in severe asthma. Am. J. Respir. Crit Care Med. 174 (2), 134-141 341. Wang, X. Nelson, A. Weiler, Z. M. Patil, A. Sato, T. Kanaji, N. Nakanishi, M. Michalski, J. Farid, M. Basma, H. Levan, T. D. Miller-Larsson, A. Wieslander, E. Muller, K. C. Holz, O. Magnussen, H. Rabe, K. F. Liu, X. Rennard, S. I. (2013) Anti-inflammatory effects of budesonide in human lung fibroblast are independent of histone deacetylase 2. J. Inflamm. Res. 6, 109-119 342. Hamid, Q. A. Wenzel, S. E. Hauk, P. J. Tsicopoulos, A. Wallaert, B. Lafitte, J. J. Chrousos, G. P. Szefler, S. J. Leung, D. Y. (1999) Increased glucocorticoid receptor beta in airway cells of glucocorticoid-insensitive asthma. Am. J. Respir. Crit Care Med. 159 (5 Pt 1), 1600-1604 343. Leung, D. Y. Hamid, Q. Vottero, A. Szefler, S. J. Surs, W. Minshall, E. Chrousos, G. P. Klemm, D. J. (1997) Association of glucocorticoid insensitivity with increased expression of glucocorticoid receptor beta. J. Exp. Med. 186 (9), 1567-1574 344. Lieber, M. Smith, B. Szakal, A. Nelson-Rees, W. Todaro, G. (1976) A continuous tumor- cell line from a human lung carcinoma with properties of type II alveolar epithelial cells. Int. J. Cancer 17 (1), 62-70 345. Bueno, O. F. De Windt, L. J. Lim, H. W. Tymitz, K. M. Witt, S. A. Kimball, T. R. Molkentin, J. D. (2001) The dual-specificity phosphatase MKP-1 limits the cardiac hypertrophic response in vitro and in vivo. Circ. Res. 88 (1), 88-96 346. Elferink, C. J. Reiners, J. J., Jr. (1996) Quantitative RT-PCR on CYP1A1 heterogeneous nuclear RNA: a surrogate for the in vitro transcription run-on assay. Biotechniques 20 (3), 470-477

202

347. Lipson, K. E. Baserga, R. (1989) Transcriptional activity of the human thymidine kinase gene determined by a method using the polymerase chain reaction and an intron-specific probe. Proc. Natl. Acad. Sci. U. S. A 86 (24), 9774-9777 348. White, R. J. (2011) Transcription by RNA polymerase III: more complex than we thought. Nat. Rev. Genet. 12 (7), 459-463 349. Mosmann, T. (1983) Rapid colorimetric assay for cellular growth and survival: application to proliferation and cytotoxicity assays. J. Immunol. Methods 65 (1-2), 55-63 350. Lasa, M. Abraham, S. M. Boucheron, C. Saklatvala, J. Clark, A. R. (2002) Dexamethasone causes sustained expression of mitogen-activated protein kinase (MAPK) phosphatase 1 and phosphatase-mediated inhibition of MAPK p38. Mol. Cell Biol. 22 (22), 7802-7811 351. Chi, H. Barry, S. P. Roth, R. J. Wu, J. J. Jones, E. A. Bennett, A. M. Flavell, R. A. (2006) Dynamic regulation of pro- and anti-inflammatory cytokines by MAPK phosphatase 1 (MKP-1) in innate immune responses. Proc. Natl. Acad. Sci. U. S. A 103 (7), 2274-2279 352. Hammer, M. Mages, J. Dietrich, H. Servatius, A. Howells, N. Cato, A. C. Lang, R. (2006) Dual specificity phosphatase 1 (DUSP1) regulates a subset of LPS-induced genes and protects mice from lethal endotoxin shock. J. Exp. Med. 203 (1), 15-20 353. Salojin, K. V. Owusu, I. B. Millerchip, K. A. Potter, M. Platt, K. A. Oravecz, T. (2006) Essential role of MAPK phosphatase-1 in the negative control of innate immune responses. J. Immunol. 176 (3), 1899-1907 354. Dauletbaev, N. Eklove, D. Mawji, N. Iskandar, M. Di, M. S. Gallouzi, I. E. Lands, L. C. (2011) Down-regulation of cytokine-induced interleukin-8 requires inhibition of p38 mitogen-activated protein kinase (MAPK) via MAPK phosphatase 1-dependent and - independent mechanisms. J. Biol. Chem. 286 (18), 15998-16007 355. Issa, R. Xie, S. Khorasani, N. Sukkar, M. Adcock, I. M. Lee, K. Y. Chung, K. F. (2007) Corticosteroid inhibition of growth-related oncogene protein-alpha via mitogen-activated kinase phosphatase-1 in airway smooth muscle cells. J. Immunol. 178 (11), 7366-7375 356. Quante, T. Ng, Y. C. Ramsay, E. E. Henness, S. Allen, J. C. Parmentier, J. Ge, Q. Ammit, A. J. (2008) Corticosteroids reduce IL-6 in ASM cells via up-regulation of MKP-1. Am. J. Respir. Cell Mol. Biol. 39 (2), 208-217 357. Szczepankiewicz, B. G. Kosogof, C. Nelson, L. T. Liu, G. Liu, B. Zhao, H. Serby, M. D. Xin, Z. Liu, M. Gum, R. J. Haasch, D. L. Wang, S. Clampit, J. E. Johnson, E. F. Lubben, T. H. Stashko, M. A. Olejniczak, E. T. Sun, C. Dorwin, S. A. Haskins, K. bad-Zapatero, C. Fry, E. H. Hutchins, C. W. Sham, H. L. Rondinone, C. M. Trevillyan, J. M. (2006) Aminopyridine-based c-Jun N-terminal kinase inhibitors with cellular activity and minimal cross-kinase activity. J. Med. Chem. 49 (12), 3563-3580 358. Holden, N. S. Catley, M. C. Cambridge, L. M. Barnes, P. J. Newton, R. (2004) ICAM-1 expression is highly NF-kappaB-dependent in A549 cells. No role for ERK and p38 MAPK. Eur. J. Biochem. 271 (4), 785-791 359. Newton, R. Cambridge, L. Hart, L. A. Stevens, D. A. Lindsay, M. A. Barnes, P. J. (2000) The MAP kinase inhibitors, PD098059, UO126 and SB203580, inhibit IL-1beta-dependent PGE(2) release via mechanistically distinct processes. Br. J. Pharmacol. 130 (6), 1353- 1361

203

360. Byon, J. C. Dadke, S. S. Rulli, S. Kusari, A. B. Kusari, J. (2001) Insulin regulates MAP kinase phosphatase-1 induction in Hirc B cells via activation of both extracellular signal- regulated kinase (ERK) and c-Jun-N-terminal kinase (JNK). Mol. Cell Biochem. 218 (1-2), 131-138 361. Junttila, M. R. Li, S. P. Westermarck, J. (2008) Phosphatase-mediated crosstalk between MAPK signaling pathways in the regulation of cell survival. FASEB J. 22 (4), 954-965 362. Arthur, J. S. Ley, S. C. (2013) Mitogen-activated protein kinases in innate immunity. Nat. Rev. Immunol. 13 (9), 679-692 363. Korhonen, R. Moilanen, E. (2014) Mitogen-activated protein kinase phosphatase 1 as an inflammatory factor and drug target. Basic Clin. Pharmacol. Toxicol. 114 (1), 24-36 364. Turpeinen, T. Nieminen, R. Moilanen, E. Korhonen, R. (2010) Mitogen-activated protein kinase phosphatase-1 negatively regulates the expression of interleukin-6, interleukin-8, and cyclooxygenase-2 in A549 human lung epithelial cells. J. Pharmacol. Exp. Ther. 333 (1), 310-318 365. Brooks, S. A. Blackshear, P. J. (2013) Tristetraprolin (TTP): interactions with mRNA and proteins, and current thoughts on mechanisms of action. Biochim. Biophys. Acta 1829 (6- 7), 666-679 366. King, E. M. Kaur, M. Gong, W. Rider, C. F. Holden, N. S. Newton, R. (2009) Regulation of tristetraprolin expression by interleukin-1beta and dexamethasone in human pulmonary epithelial cells: roles for nuclear factor-kappaB and p38 mitogen-activated protein kinase. J. Pharmacol. Exp. Ther. 330 (2), 575-585 367. Jalonen, U. Lahti, A. Korhonen, R. Kankaanranta, H. Moilanen, E. (2005) Inhibition of tristetraprolin expression by dexamethasone in activated macrophages. Biochem. Pharmacol. 69 (5), 733-740 368. Smoak, K. Cidlowski, J. A. (2006) Glucocorticoids regulate tristetraprolin synthesis and posttranscriptionally regulate tumor necrosis factor alpha inflammatory signaling. Mol. Cell Biol. 26 (23), 9126-9135 369. Ishmael, F. T. Fang, X. Galdiero, M. R. Atasoy, U. Rigby, W. F. Gorospe, M. Cheadle, C. Stellato, C. (2008) Role of the RNA-Binding Protein Tristetraprolin in Glucocorticoid- Mediated Gene Regulation. J. Immunol. 180 (12), 8342-8353 370. Ayroldi, E. Riccardi, C. (2009) Glucocorticoid-induced leucine zipper (GILZ): a new important mediator of glucocorticoid action. FASEB J. 23 (11), 3649-3658 371. Chang, T. S. Kim, M. J. Ryoo, K. Park, J. Eom, S. J. Shim, J. Nakayama, K. I. Nakayama, K. Tomita, M. Takahashi, K. Lee, M. J. Choi, E. J. (2003) p57KIP2 modulates stress- activated signaling by inhibiting c-Jun NH2-terminal kinase/stress-activated protein Kinase. J. Biol. Chem. 278 (48), 48092-48098 372. Kelly, M. M. King, E. M. Rider, C. F. Gwozd, C. Holden, N. S. Eddleston, J. Zuraw, B. Leigh, R. O'Byrne, P. M. Newton, R. (2012) Corticosteroid-induced gene expression in allergen-challenged asthmatic subjects taking inhaled budesonide. Br. J. Pharmacol. 165 (6), 1737-1747 373. Rider, C. F. King, E. M. Holden, N. S. Giembycz, M. A. Newton, R. (2011) Inflammatory stimuli inhibit glucocorticoid-dependent transactivation in human pulmonary epithelial

204

cells: rescue by long-acting beta2-adrenoceptor agonists. J. Pharmacol. Exp. Ther. 338 (3), 860-869 374. Falvo, J. V. Tsytsykova, A. V. Goldfeld, A. E. (2010) Transcriptional control of the TNF gene. Curr. Dir. Autoimmun. 11, 27-60 375. Carballo, E. Lai, W. S. Blackshear, P. J. (1998) Feedback inhibition of macrophage tumor necrosis factor-alpha production by tristetraprolin. Science 281 (5379), 1001-1005 376. Dean, J. L. Sarsfield, S. J. Tsounakou, E. Saklatvala, J. (2003) p38 Mitogen-activated protein kinase stabilizes mRNAs that contain cyclooxygenase-2 and tumor necrosis factor AU-rich elements by inhibiting deadenylation. J. Biol. Chem. 278 (41), 39470-39476 377. Lai, W. S. Carballo, E. Strum, J. R. Kennington, E. A. Phillips, R. S. Blackshear, P. J. (1999) Evidence that tristetraprolin binds to AU-rich elements and promotes the deadenylation and destabilization of tumor necrosis factor alpha mRNA. Mol. Cell Biol. 19 (6), 4311-4323 378. Darnell, J. E., Jr. (2013) Reflections on the history of pre-mRNA processing and highlights of current knowledge: a unified picture. RNA. 19 (4), 443-460 379. McGeehan, G. M. Becherer, J. D. Bast, R. C., Jr. Boyer, C. M. Champion, B. Connolly, K. M. Conway, J. G. Furdon, P. Karp, S. Kidao, S. . (1994) Regulation of tumour necrosis factor-alpha processing by a metalloproteinase inhibitor. Nature 370 (6490), 558-561 380. Ronkina, N. Menon, M. B. Schwermann, J. Tiedje, C. Hitti, E. Kotlyarov, A. Gaestel, M. (2010) MAPKAP kinases MK2 and MK3 in inflammation: complex regulation of TNF biosynthesis via expression and phosphorylation of tristetraprolin. Biochem. Pharmacol. 80 (12), 1915-1920 381. Khair, O. A. Davies, R. J. Devalia, J. L. (1996) Bacterial-induced release of inflammatory mediators by bronchial epithelial cells. Eur. Respir. J. 9 (9), 1913-1922 382. Shah, S. King, E. M. Chandrasekhar, A. Newton, R. (2014) Roles for the mitogen- activated protein kinase (MAPK) phosphatase, DUSP1, in feedback control of inflammatory gene expression and repression by dexamethasone. J. Biol. Chem. 289 (19), 13667-13679 383. Deleault, K. M. Skinner, S. J. Brooks, S. A. (2008) Tristetraprolin regulates TNF TNF- alpha mRNA stability via a proteasome dependent mechanism involving the combined action of the ERK and p38 pathways. Mol. Immunol. 45 (1), 13-24 384. Brook, M. Sully, G. Clark, A. R. Saklatvala, J. (2000) Regulation of tumour necrosis factor alpha mRNA stability by the mitogen-activated protein kinase p38 signalling cascade. FEBS Lett. 483 (1), 57-61 385. Dean, J. L. Brook, M. Clark, A. R. Saklatvala, J. (1999) p38 mitogen-activated protein kinase regulates cyclooxygenase-2 mRNA stability and transcription in lipopolysaccharide-treated human monocytes. J. Biol. Chem. 274 (1), 264-269 386. Dean, J. L. Sully, G. Clark, A. R. Saklatvala, J. (2004) The involvement of AU-rich element-binding proteins in p38 mitogen-activated protein kinase pathway-mediated mRNA stabilisation. Cell Signal. 16 (10), 1113-1121 387. Tudor, C. Marchese, F. P. Hitti, E. Aubareda, A. Rawlinson, L. Gaestel, M. Blackshear, P. J. Clark, A. R. Saklatvala, J. Dean, J. L. (2009) The p38 MAPK pathway inhibits

205

tristetraprolin-directed decay of interleukin-10 and pro-inflammatory mediator mRNAs in murine macrophages. FEBS Lett. 583 (12), 1933-1938 388. Lu, J. Y. Sadri, N. Schneider, R. J. (2006) Endotoxic shock in AUF1 knockout mice mediated by failure to degrade proinflammatory cytokine mRNAs. Genes Dev. 20 (22), 3174-3184 389. Katsanou, V. Papadaki, O. Milatos, S. Blackshear, P. J. Anderson, P. Kollias, G. Kontoyiannis, D. L. (2005) HuR as a negative posttranscriptional modulator in inflammation. Mol. Cell 19 (6), 777-789 390. Clark, A. R. Belvisi, M. G. (2012) Maps and legends: the quest for dissociated ligands of the glucocorticoid receptor. Pharmacol. Ther. 134 (1), 54-67 391. Jalonen, U. Lahti, A. Korhonen, R. Kankaanranta, H. Moilanen, E. (2005) Inhibition of tristetraprolin expression by dexamethasone in activated macrophages. Biochem. Pharmacol. 69 (5), 733-740 392. Newton, R. (2014) Anti-inflammatory glucocorticoids: changing concepts. Eur. J. Pharmacol. 724, 231-236 393. Baumgartner, R. A. Deramo, V. A. Beaven, M. A. (1996) Constitutive and inducible mechanisms for synthesis and release of cytokines in immune cell lines. J. Immunol. 157 (9), 4087-4093 394. Bissonnette, E. Y. Enciso, J. A. Befus, A. D. (1995) Inhibition of tumour necrosis factor- alpha (TNF-alpha) release from mast cells by the anti-inflammatory drugs, sodium cromoglycate and nedocromil sodium. Clin. Exp. Immunol. 102 (1), 78-84 395. Lippert, U. Moller, A. Welker, P. Artuc, M. Henz, B. M. (2000) Inhibition of cytokine secretion from human leukemic mast cells and basophils by H1- and H2-receptor antagonists. Exp. Dermatol. 9 (2), 118-124 396. Williams, C. M. Coleman, J. W. (1995) Induced expression of mRNA for IL-5, IL-6, TNF- alpha, MIP-2 and IFN-gamma in immunologically activated rat peritoneal mast cells: inhibition by dexamethasone and cyclosporin A. Immunology 86 (2), 244-249 397. Zhao, Y. Leung, P. C. Woo, K. S. Chen, G. G. Wong, Y. O. Liu, S. X. van Hasselt, C. A. (2004) Inhibitory effects of budesonide, desloratadine and dexamethasone on cytokine release from human mast cell line (HMC-1). Inflamm. Res. 53 (12), 664-669 398. Holden, N. S. Gong, W. King, E. M. Kaur, M. Giembycz, M. A. Newton, R. (2007) Potentiation of NF-kappaB-dependent transcription and inflammatory mediator release by histamine in human airway epithelial cells. Br. J. Pharmacol. 152 (6), 891-902 399. Grell, M. Douni, E. Wajant, H. Lohden, M. Clauss, M. Maxeiner, B. Georgopoulos, S. Lesslauer, W. Kollias, G. Pfizenmaier, K. Scheurich, P. (1995) The transmembrane form of tumor necrosis factor is the prime activating ligand of the 80 kDa tumor necrosis factor receptor. Cell 83 (5), 793-802 400. Horiuchi, T. Mitoma, H. Harashima, S. Tsukamoto, H. Shimoda, T. (2010) Transmembrane TNF-alpha: structure, function and interaction with anti-TNF agents. Rheumatology. (Oxford) 49 (7), 1215-1228

206

401. Solomon, K. A. Covington, M. B. DeCicco, C. P. Newton, R. C. (1997) The fate of pro- TNF-alpha following inhibition of metalloprotease-dependent processing to soluble TNF- alpha in human monocytes. J. Immunol. 159 (9), 4524-4531 402. Crawford, E. K. Ensor, J. E. Kalvakolanu, I. Hasday, J. D. (1997) The role of 3' poly(A) tail metabolism in tumor necrosis factor-alpha regulation. J. Biol. Chem. 272 (34), 21120- 21127 403. Swantek, J. L. Cobb, M. H. Geppert, T. D. (1997) Jun N-terminal kinase/stress-activated protein kinase (JNK/SAPK) is required for lipopolysaccharide stimulation of tumor necrosis factor alpha (TNF-alpha) translation: glucocorticoids inhibit TNF-alpha translation by blocking JNK/SAPK. Mol. Cell Biol. 17 (11), 6274-6282 404. Prabhala, P. Bunge, K. Rahman, M. M. Ge, Q. Clark, A. R. Ammit, A. J. (2015) Temporal regulation of cytokine mRNA expression by tristetraprolin: dynamic control by p38 MAPK and MKP-1. Am. J. Physiol Lung Cell Mol. Physiol 308 (9), L973-L980 405. Smallie, T. Ross, E. A. Ammit, A. J. Cunliffe, H. E. Tang, T. Rosner, D. R. Ridley, M. L. Buckley, C. D. Saklatvala, J. Dean, J. L. Clark, A. R. (2015) Dual-Specificity Phosphatase 1 and Tristetraprolin Cooperate To Regulate Macrophage Responses to Lipopolysaccharide. J. Immunol. 195 (1), 277-288 406. Clark, A. R. Dean, J. L. Saklatvala, J. (2009) The p38 MAPK pathway mediates both antiinflammatory and proinflammatory processes: comment on the article by Damjanov and the editorial by Genovese. Arthritis Rheum. 60 (11), 3513-3514 407. Chapin, W. J. Lenkala, D. Mai, Y. Mao, Y. White, S. R. Huang, R. S. (2015) Peripheral blood IRF1 expression as a marker for glucocorticoid sensitivity. Pharmacogenet. Genomics 25 (3), 126-133 408. Nakagawa, K. Yokosawa, H. (2000) Degradation of transcription factor IRF-1 by the ubiquitin-proteasome pathway. The C-terminal region governs the protein stability. Eur. J. Biochem. 267 (6), 1680-1686 409. Watanabe, N. Sakakibara, J. Hovanessian, A. G. Taniguchi, T. Fujita, T. (1991) Activation of IFN-beta element by IRF-1 requires a posttranslational event in addition to IRF-1 synthesis. Nucleic Acids Res. 19 (16), 4421-4428 410. Stellato, C. (2007) Glucocorticoid actions on airway epithelial responses in immunity: functional outcomes and molecular targets. J. Allergy Clin. Immunol. 120 (6), 1247-1263 411. Zhang, N. Truong-Tran, Q. A. Tancowny, B. Harris, K. E. Schleimer, R. P. (2007) Glucocorticoids enhance or spare innate immunity: effects in airway epithelium are mediated by CCAAT/enhancer binding proteins. J. Immunol. 179 (1), 578-589 412. Hui, L. Bakiri, L. Mairhorfer, A. Schweifer, N. Haslinger, C. Kenner, L. Komnenovic, V. Scheuch, H. Beug, H. Wagner, E. F. (2007) p38alpha suppresses normal and cancer cell proliferation by antagonizing the JNK-c-Jun pathway. Nat. Genet. 39 (6), 741-749 413. Jeffrey, K. L. Brummer, T. Rolph, M. S. Liu, S. M. Callejas, N. A. Grumont, R. J. Gillieron, C. Mackay, F. Grey, S. Camps, M. Rommel, C. Gerondakis, S. D. Mackay, C. R. (2006) Positive regulation of immune cell function and inflammatory responses by phosphatase PAC-1. Nat. Immunol. 7 (3), 274-283

207

414. Whitmarsh, A. J. (2007) Regulation of gene transcription by mitogen-activated protein kinase signaling pathways. Biochim. Biophys. Acta 1773 (8), 1285-1298 415. Zhou, X. Richon, V. M. Wang, A. H. Yang, X. J. Rifkind, R. A. Marks, P. A. (2000) Histone deacetylase 4 associates with extracellular signal-regulated kinases 1 and 2, and its cellular localization is regulated by oncogenic Ras. Proc. Natl. Acad. Sci. U. S. A 97 (26), 14329-14333 416. Kim, T. K. Maniatis, T. (1996) Regulation of interferon-gamma-activated STAT1 by the ubiquitin-proteasome pathway. Science 273 (5282), 1717-1719 417. Maki, C. G. Huibregtse, J. M. Howley, P. M. (1996) In vivo ubiquitination and proteasome-mediated degradation of p53(1). Cancer Res. 56 (11), 2649-2654 418. Marti, A. Wirbelauer, C. Scheffner, M. Krek, W. (1999) Interaction between ubiquitin- protein ligase SCFSKP2 and -1 underlies the regulation of E2F-1 degradation. Nat. Cell Biol. 1 (1), 14-19 419. Salghetti, S. E. Kim, S. Y. Tansey, W. P. (1999) Destruction of by ubiquitin-mediated proteolysis: cancer-associated and transforming mutations stabilize Myc. EMBO J. 18 (3), 717-726 420. Stancovski, I. Gonen, H. Orian, A. Schwartz, A. L. Ciechanover, A. (1995) Degradation of the proto-oncogene product c-Fos by the ubiquitin proteolytic system in vivo and in vitro: identification and characterization of the conjugating enzymes. Mol. Cell Biol. 15 (12), 7106-7116 421. Treier, M. Staszewski, L. M. Bohmann, D. (1994) Ubiquitin-dependent c-Jun degradation in vivo is mediated by the delta domain. Cell 78 (5), 787-798 422. Laine, A. Ronai, Z. (2005) Ubiquitin chains in the ladder of MAPK signaling. Sci. STKE. 2005 (281), re5 423. Huang, Y. Krein, P. M. Muruve, D. A. Winston, B. W. (2002) Complement factor B gene regulation: synergistic effects of TNF-alpha and IFN-gamma in macrophages. J. Immunol. 169 (5), 2627-2635 424. Shi, L. Perin, J. C. Leipzig, J. Zhang, Z. Sullivan, K. E. (2011) Genome-wide analysis of interferon regulatory factor I binding in primary human monocytes. Gene 487 (1), 21-28 425. Miyamoto, N. G. Medberry, P. S. Hesselgesser, J. Boehlk, S. Nelson, P. J. Krensky, A. M. Perez, H. D. (2000) Interleukin-1beta induction of the chemokine RANTES promoter in the human astrocytoma line CH235 requires both constitutive and inducible transcription factors. J. Neuroimmunol. 105 (1), 78-90 426. Renauld, J. C. (2001) New insights into the role of cytokines in asthma. J. Clin. Pathol. 54 (8), 577-589 427. Carter, A. B. Knudtson, K. L. Monick, M. M. Hunninghake, G. W. (1999) The p38 mitogen-activated protein kinase is required for NF-kappaB-dependent gene expression. The role of TATA-binding protein (TBP). J. Biol. Chem. 274 (43), 30858-30863 428. Carter, A. B. Hunninghake, G. W. (2000) A constitutive active MEK --> ERK pathway negatively regulates NF-kappa B-dependent gene expression by modulating TATA- binding protein phosphorylation. J. Biol. Chem. 275 (36), 27858-27864

208

429. Bain, J. Plater, L. Elliott, M. Shpiro, N. Hastie, C. J. McLauchlan, H. Klevernic, I. Arthur, J. S. Alessi, D. R. Cohen, P. (2007) The selectivity of protein kinase inhibitors: a further update. Biochem. J. 408 (3), 297-315 430. Davies, S. P. Reddy, H. Caivano, M. Cohen, P. (2000) Specificity and mechanism of action of some commonly used protein kinase inhibitors. Biochem. J. 351 (Pt 1), 95-105 431. Pawliczak, R. Logun, C. Madara, P. Barb, J. Suffredini, A. F. Munson, P. J. Danner, R. L. Shelhamer, J. H. (2005) Influence of IFN-gamma on gene expression in normal human bronchial epithelial cells: modulation of IFN-gamma effects by dexamethasone. Physiol Genomics 23 (1), 28-45 432. van de Garde, M. D. Martinez, F. O. Melgert, B. N. Hylkema, M. N. Jonkers, R. E. Hamann, J. (2014) Chronic exposure to glucocorticoids shapes gene expression and modulates innate and adaptive activation pathways in macrophages with distinct changes in leukocyte attraction. J. Immunol. 192 (3), 1196-1208 433. Griffith, J. W. Sokol, C. L. Luster, A. D. (2014) Chemokines and chemokine receptors: positioning cells for host defense and immunity. Annu. Rev. Immunol. 32, 659-702 434. Wark, P. A. Bucchieri, F. Johnston, S. L. Gibson, P. G. Hamilton, L. Mimica, J. Zummo, G. Holgate, S. T. Attia, J. Thakkinstian, A. Davies, D. E. (2007) IFN-gamma-induced protein 10 is a novel biomarker of rhinovirus-induced asthma exacerbations. J. Allergy Clin. Immunol. 120 (3), 586-593 435. Karin, M. (2005) Inflammation-activated protein kinases as targets for drug development. Proc. Am. Thorac. Soc. 2 (4), 386-390 436. Sridhar, R. Hanson-Painton, O. Cooper, D. R. (2000) Protein kinases as therapeutic targets. Pharm. Res. 17 (11), 1345-1353 437. Adcock, I. M. Caramori, G. (2004) Kinase targets and inhibitors for the treatment of airway inflammatory diseases: the next generation of drugs for severe asthma and COPD? BioDrugs. 18 (3), 167-180 438. Caunt, C. J. Keyse, S. M. (2013) Dual-specificity MAP kinase phosphatases (MKPs): shaping the outcome of MAP kinase signalling. FEBS J. 280 (2), 489-504 439. Bluthgen, N. Legewie, S. (2008) Systems analysis of MAPK signal transduction. Essays Biochem. 45, 95-107 440. Bluthgen, N. Legewie, S. Kielbasa, S. M. Schramme, A. Tchernitsa, O. Keil, J. Solf, A. Vingron, M. Schafer, R. Herzel, H. Sers, C. (2009) A systems biological approach suggests that transcriptional feedback regulation by dual-specificity phosphatase 6 shapes extracellular signal-related kinase activity in RAS-transformed fibroblasts. FEBS J. 276 (4), 1024-1035 441. Kholodenko, B. N. (2000) Negative feedback and ultrasensitivity can bring about oscillations in the mitogen-activated protein kinase cascades. Eur. J. Biochem. 267 (6), 1583-1588 442. Chrestensen, C. A. Schroeder, M. J. Shabanowitz, J. Hunt, D. F. Pelo, J. W. Worthington, M. T. Sturgill, T. W. (2004) MAPKAP kinase 2 phosphorylates tristetraprolin on in vivo sites including Ser178, a site required for 14-3-3 binding. J. Biol. Chem. 279 (11), 10176- 10184

209

443. Hitti, E. Iakovleva, T. Brook, M. Deppenmeier, S. Gruber, A. D. Radzioch, D. Clark, A. R. Blackshear, P. J. Kotlyarov, A. Gaestel, M. (2006) Mitogen-activated protein kinase- activated protein kinase 2 regulates tumor necrosis factor mRNA stability and translation mainly by altering tristetraprolin expression, stability, and binding to adenine/uridine-rich element. Mol. Cell Biol. 26 (6), 2399-2407 444. Rincon, M. Enslen, H. Raingeaud, J. Recht, M. Zapton, T. Su, M. S. Penix, L. A. Davis, R. J. Flavell, R. A. (1998) Interferon-gamma expression by Th1 effector T cells mediated by the p38 MAP kinase signaling pathway. EMBO J. 17 (10), 2817-2829 445. Hoebe, K. Janssen, E. Beutler, B. (2004) The interface between innate and adaptive immunity. Nat. Immunol. 5 (10), 971-974 446. Salojin, K. Oravecz, T. (2007) Regulation of innate immunity by MAPK dual-specificity phosphatases: knockout models reveal new tricks of old genes. J. Leukoc. Biol. 81 (4), 860- 869 447. Sokol, C. L. Luster, A. D. (2015) The chemokine system in innate immunity. Cold Spring Harb. Perspect. Biol. 7 (5) 448. Lack, G. Bradley, K. L. Hamelmann, E. Renz, H. Loader, J. Leung, D. Y. Larsen, G. Gelfand, E. W. (1996) Nebulized IFN-gamma inhibits the development of secondary allergic responses in mice. J. Immunol. 157 (4), 1432-1439 449. Sakai, A. Han, J. Cato, A. C. Akira, S. Li, J. D. (2004) Glucocorticoids synergize with IL- 1beta to induce TLR2 expression via MAP Kinase Phosphatase-1-dependent dual Inhibition of MAPK JNK and p38 in epithelial cells. BMC. Mol. Biol. 5, 2 450. Ehrchen, J. Steinmuller, L. Barczyk, K. Tenbrock, K. Nacken, W. Eisenacher, M. Nordhues, U. Sorg, C. Sunderkotter, C. Roth, J. (2007) Glucocorticoids induce differentiation of a specifically activated, anti-inflammatory subtype of human monocytes. Blood 109 (3), 1265-1274 451. Galon, J. Franchimont, D. Hiroi, N. Frey, G. Boettner, A. Ehrhart-Bornstein, M. O'Shea, J. J. Chrousos, G. P. Bornstein, S. R. (2002) Gene profiling reveals unknown enhancing and suppressive actions of glucocorticoids on immune cells. FASEB J. 16 (1), 61-71 452. Hermoso, M. A. Matsuguchi, T. Smoak, K. Cidlowski, J. A. (2004) Glucocorticoids and tumor necrosis factor alpha cooperatively regulate toll-like receptor 2 gene expression. Mol. Cell Biol. 24 (11), 4743-4756 453. Homma, T. Kato, A. Hashimoto, N. Batchelor, J. Yoshikawa, M. Imai, S. Wakiguchi, H. Saito, H. Matsumoto, K. (2004) Corticosteroid and cytokines synergistically enhance toll- like receptor 2 expression in respiratory epithelial cells. Am. J. Respir. Cell Mol. Biol. 31 (4), 463-469 454. Imasato, A. sbois-Mouthon, C. Han, J. Kai, H. Cato, A. C. Akira, S. Li, J. D. (2002) Inhibition of p38 MAPK by glucocorticoids via induction of MAPK phosphatase-1 enhances nontypeable Haemophilus influenzae-induced expression of toll-like receptor 2. J. Biol. Chem. 277 (49), 47444-47450 455. Shuto, T. Imasato, A. Jono, H. Sakai, A. Xu, H. Watanabe, T. Rixter, D. D. Kai, H. Andalibi, A. Linthicum, F. Guan, Y. L. Han, J. Cato, A. C. Lim, D. J. Akira, S. Li, J. D. (2002) Glucocorticoids synergistically enhance nontypeable Haemophilus influenzae-

210

induced Toll-like receptor 2 expression via a negative cross-talk with p38 MAP kinase. J. Biol. Chem. 277 (19), 17263-17270 456. Newton, R. Holden, N. S. (2007) Separating transrepression and transactivation: a distressing divorce for the glucocorticoid receptor? Mol. Pharmacol. 72 (4), 799-809 457. Clark, A. R. Belvisi, M. G. (2012) Maps and legends: the quest for dissociated ligands of the glucocorticoid receptor. Pharmacol. Ther. 134 (1), 54-67 458. Clark, A. R. Martins, J. R. Tchen, C. R. (2008) Role of dual specificity phosphatases in biological responses to glucocorticoids. J. Biol. Chem. 283 (38), 25765-25769 459. Korhonen, R. Moilanen, E. (2013) MAP Kinase Phosphatase-1 as an Inflammatory Factor and Drug Target. Basic Clin. Pharmacol. Toxicol. 460. King, E. M. Chivers, J. E. Rider, C. F. Minnich, A. Giembycz, M. A. Newton, R. (2013) Glucocorticoid Repression of Inflammatory Gene Expression Shows Differential Responsiveness by Transactivation- and Transrepression-Dependent Mechanisms. PLoS ONE 8 (1), e53936 461. Auphan, N. DiDonato, J. A. Rosette, C. Helmberg, A. Karin, M. (1995) Immunosuppression by glucocorticoids: inhibition of NF-kappa B activity through induction of I kappa B synthesis. Science 270 (5234), 286-290 462. Ayroldi, E. Migliorati, G. Bruscoli, S. Marchetti, C. Zollo, O. Cannarile, L. D'Adamio, F. Riccardi, C. (2001) Modulation of T-cell activation by the glucocorticoid-induced leucine zipper factor via inhibition of nuclear factor kappaB. Blood 98 (3), 743-753 463. Eddleston, J. Herschbach, J. Wagelie-Steffen, A. L. Christiansen, S. C. Zuraw, B. L. (2007) The anti-inflammatory effect of glucocorticoids is mediated by glucocorticoid- induced leucine zipper in epithelial cells. J. Allergy Clin. Immunol. 119 (1), 115-122 464. Sadler, A. J. Suliman, B. A. Yu, L. Yuan, X. Wang, D. Irving, A. T. Sarvestani, S. T. Banerjee, A. Mansell, A. S. Liu, J. P. Gerondakis, S. Williams, B. R. Xu, D. (2015) The acetyltransferase HAT1 moderates the NF-kappaB response by regulating the transcription factor PLZF. Nat. Commun. 6, 6795 465. Sadler, A. J. Rossello, F. J. Yu, L. Deane, J. A. Yuan, X. Wang, D. Irving, A. T. Kaparakis- Liaskos, M. Gantier, M. P. Ying, H. Yim, H. C. Hartland, E. L. Notini, A. J. de, B. S. White, S. J. Mansell, A. Liu, J. P. Watkins, D. N. Gerondakis, S. Williams, B. R. Xu, D. (2015) BTB-ZF transcriptional regulator PLZF modifies chromatin to restrain inflammatory signaling programs. Proc. Natl. Acad. Sci. U. S. A 112 (5), 1535-1540 466. Liang, J. Lei, T. Song, Y. Yanes, N. Qi, Y. Fu, M. (2009) RNA-destabilizing factor tristetraprolin negatively regulates NF-kappaB signaling. J. Biol. Chem. 284 (43), 29383- 29390 467. Altonsy, M. O. Sasse, S. K. Phang, T. L. Gerber, A. N. (2014) Context-dependent cooperation between nuclear factor kappaB (NF-kappaB) and the glucocorticoid receptor at a TNFAIP3 intronic enhancer: a mechanism to maintain negative feedback control of inflammation. J. Biol. Chem. 289 (12), 8231-8239 468. Miyata, M. Lee, J. Y. Susuki-Miyata, S. Wang, W. Y. Xu, H. Kai, H. Kobayashi, K. S. Flavell, R. A. Li, J. D. (2015) Glucocorticoids suppress inflammation via the upregulation of negative regulator IRAK-M. Nat. Commun. 6, 6062

211

469. Manetsch, M. Ramsay, E. E. King, E. M. Seidel, P. Che, W. Ge, Q. Hibbs, D. E. Newton, R. Ammit, A. J. (2012) Corticosteroids and beta(2)-agonists upregulate mitogen-activated protein kinase phosphatase 1: in vitro mechanisms. Br. J. Pharmacol. 166 (7), 2049-2059 470. Rao, N. A. McCalman, M. T. Moulos, P. Francoijs, K. J. Chatziioannou, A. Kolisis, F. N. Alexis, M. N. Mitsiou, D. J. Stunnenberg, H. G. (2011) Coactivation of GR and NFKB alters the repertoire of their binding sites and target genes. Genome Res. 21 (9), 1404-1416 471. Wada, T. Takagi, T. Yamaguchi, Y. Ferdous, A. Imai, T. Hirose, S. Sugimoto, S. Yano, K. Hartzog, G. A. Winston, F. Buratowski, S. Handa, H. (1998) DSIF, a novel transcription elongation factor that regulates RNA polymerase II processivity, is composed of human Spt4 and Spt5 homologs. Genes Dev. 12 (3), 343-356 472. Yamaguchi, Y. Takagi, T. Wada, T. Yano, K. Furuya, A. Sugimoto, S. Hasegawa, J. Handa, H. (1999) NELF, a multisubunit complex containing RD, cooperates with DSIF to repress RNA polymerase II elongation. Cell 97 (1), 41-51 473. Ping, Y. H. Rana, T. M. (2001) DSIF and NELF interact with RNA polymerase II elongation complex and HIV-1 Tat stimulates P-TEFb-mediated phosphorylation of RNA polymerase II and DSIF during transcription elongation. J. Biol. Chem. 276 (16), 12951- 12958 474. Nissen, R. M. Yamamoto, K. R. (2000) The glucocorticoid receptor inhibits NFkappaB by interfering with serine-2 phosphorylation of the RNA polymerase II carboxy-terminal domain. Genes Dev. 14 (18), 2314-2329 475. Luecke, H. F. Yamamoto, K. R. (2005) The glucocorticoid receptor blocks P-TEFb recruitment by NFkappaB to effect promoter-specific transcriptional repression. Genes Dev. 19 (9), 1116-1127 476. Jatakanon, A. Uasuf, C. Maziak, W. Lim, S. Chung, K. F. Barnes, P. J. (1999) Neutrophilic inflammation in severe persistent asthma. Am. J. Respir. Crit Care Med. 160 (5 Pt 1), 1532- 1539 477. Schacke, H. Docke, W. D. Asadullah, K. (2002) Mechanisms involved in the side effects of glucocorticoids. Pharmacol. Ther. 96 (1), 23-43 478. Vayssiere, B. M. Dupont, S. Choquart, A. Petit, F. Garcia, T. Marchandeau, C. Gronemeyer, H. Resche-Rigon, M. (1997) Synthetic glucocorticoids that dissociate transactivation and AP-1 transrepression exhibit antiinflammatory activity in vivo. Mol. Endocrinol. 11 (9), 1245-1255 479. Vanden Berghe, W. Francesconi, E. De, B. K. Resche-Rigon, M. Haegeman, G. (1999) Dissociated glucocorticoids with anti-inflammatory potential repress interleukin-6 gene expression by a nuclear factor-kappaB-dependent mechanism. Mol. Pharmacol. 56 (4), 797-806 480. Belvisi, M. G. Wicks, S. L. Battram, C. H. Bottoms, S. E. Redford, J. E. Woodman, P. Brown, T. J. Webber, S. E. Foster, M. L. (2001) Therapeutic benefit of a dissociated glucocorticoid and the relevance of in vitro separation of transrepression from transactivation activity. J. Immunol. 166 (3), 1975-1982

212

481. Boruk, M. Savory, J. G. Hache, R. J. (1998) AF-2-dependent potentiation of CCAAT enhancer binding protein beta-mediated transcriptional activation by glucocorticoid receptor. Mol. Endocrinol. 12 (11), 1749-1763 482. Kordula, T. Travis, J. (1996) The role of Stat and C/EBP transcription factors in the synergistic activation of rat serine protease inhibitor-3 gene by interleukin-6 and dexamethasone. Biochem. J. 313 ( Pt 3), 1019-1027 483. Diamond, M. I. Miner, J. N. Yoshinaga, S. K. Yamamoto, K. R. (1990) Transcription factor interactions: selectors of positive or negative regulation from a single DNA element. Science 249 (4974), 1266-1272 484. Hofmann, T. G. Schmitz, M. L. (2002) The promoter context determines mutual repression or synergism between NF-kappaB and the glucocorticoid receptor. Biol. Chem. 383 (12), 1947-1951 485. So, A. Y. Chaivorapol, C. Bolton, E. C. Li, H. Yamamoto, K. R. (2007) Determinants of cell- and gene-specific transcriptional regulation by the glucocorticoid receptor. PLoS. Genet. 3 (6), e94 486. Rogatsky, I. Wang, J. C. Derynck, M. K. Nonaka, D. F. Khodabakhsh, D. B. Haqq, C. M. Darimont, B. D. Garabedian, M. J. Yamamoto, K. R. (2003) Target-specific utilization of transcriptional regulatory surfaces by the glucocorticoid receptor. Proc. Natl. Acad. Sci. U. S. A 100 (24), 13845-13850 487. Reichardt, H. M. Tuckermann, J. P. Gottlicher, M. Vujic, M. Weih, F. Angel, P. Herrlich, P. Schutz, G. (2001) Repression of inflammatory responses in the absence of DNA binding by the glucocorticoid receptor. EMBO J. 20 (24), 7168-7173 488. Page, C. P. Spina, D. (2012) Selective PDE inhibitors as novel treatments for respiratory diseases. Curr. Opin. Pharmacol. 12 (3), 275-286 489. Patel, B. S. Prabhala, P. Oliver, B. G. Ammit, A. J. (2015) Inhibitors of Phosphodiesterase 4, but Not Phosphodiesterase 3, Increase beta2-Agonist-Induced Expression of Antiinflammatory Mitogen-Activated Protein Kinase Phosphatase 1 in Airway Smooth Muscle Cells. Am. J. Respir. Cell Mol. Biol. 52 (5), 634-640 490. Korhonen, R. Hommo, T. Keranen, T. Laavola, M. Hamalainen, M. Vuolteenaho, K. Lehtimaki, L. Kankaanranta, H. Moilanen, E. (2013) Attenuation of TNF production and experimentally induced inflammation by PDE4 inhibitor rolipram is mediated by MAPK phosphatase-1. Br. J. Pharmacol. 169 (7), 1525-1536 491. Giembycz, M. A. (2008) Can the anti-inflammatory potential of PDE4 inhibitors be realized: guarded optimism or wishful thinking? Br. J. Pharmacol. 155 (3), 288-290 492. Giembycz, M. A. Newton, R. (2015) Potential mechanisms to explain how LABAs and PDE4 inhibitors enhance the clinical efficacy of glucocorticoids in inflammatory lung diseases. F1000Prime. Rep. 7, 16 493. Horsch, K. de, W. H. Schuurmans, M. M. Allie-Reid, F. Cato, A. C. Cunningham, J. Burrin, J. M. Hough, F. S. Hulley, P. A. (2007) Mitogen-activated protein kinase phosphatase 1/dual specificity phosphatase 1 mediates glucocorticoid inhibition of osteoblast proliferation. Mol. Endocrinol. 21 (12), 2929-2940

213

Appendix A: Antibodies, siRNA, PCR and ChIP primers

Table A1. siRNA sequence for knockdown studies

(IRF-1 siRNA was provided by Dr. David Proud, University of Calgary, Calgary, Canada and others were supplied by Qiagen, Canada.)

siRNA Target Sequence Lamin A/C (LMNA) 5’-AACTGGACTTCCAGAAGAACA-3’ siRNA DUSP1 siRNA 1 5’-TAGCGTCAAGACATTTGCTGA-3’ DUSP1 siRNA 2 5’-CTGTACTATCCTGTAAATATA-3’ IRF1 siRNA 1 5’-GGGACAUCAACAAGGAUGCCUGUUU-3’ IRF1 siRNA 2 5’-CGGACAGCACCAGTGATCTGTACAA-3’ ZFP36 siRNA 1 5’-ACCGACGATATAATTATTATA-3’ ZFP36 siRNA 2 5’-ACGACTTTATTTATTCTAATA-3’

214

Table A2. Antibodies used for western blot analysis

List of antibodies including the working dilution of primary antibody. Antibodies were supplied by: AbD Serotec, Raleigh, NC, USA; Santa Cruz Biotechnology Inc., Santa Cruz, CA, USA; Cell

Signalling, Danvers, MA, USA; Abcam Inc., Cambridge, MA, USA.

Primary Secondary Dilution Company Cat. number Antibody

cJUN Rabbit 1:1000 Cell Signalling #9165 DUSP1 Rabbit 1:250 Santa Cruz sc-1102 GAPDH Mouse 1:40000 AbD Serotac 4699-9555 IRF1 Rabbit 1:15000 Santa Cruz sc-497 Phosho-cJun Rabbit 1:1000 Cell Signalling #9164 phospho-JNK Rabbit 1:1000 Cell Signalling #9251 (Thr183/Tyr185) phospho-p38 Rabbit 1:1000 Cell Signalling #9211 (Thr180/Tyr182) phospho-p44/42 Rabbit 1:1000 Cell Signalling #9101 (Thr202/Tyr204) PTGS2 Goat 1: 1000 Santa Cruz sc-1746 TNF Rabbit 1:1000 Abcam ab66579 ZFP36 Rabbit 1:500 Santa Cruz sc-14030

215

Table A3. Primers used for PCR analysis

Forward (F) and reverse (R) primer sequences (5’-3’) are shown in addition to the accession number for each gene. For genes with more than one splice variant, primers were designed to pick up all variants, with the exception of IFIT3 for which two sets of primers were designed. All primers were designed using Primer Express software (Applied Biosystems) and were synthesised by the DNA synthesis lab at the University of Calgary.

Official Human Genome Organization (HUGO) gene nomenclature committee gene symbols have been used for all genes and gene products.

Gene Symbol Accession Primer Sequences Number APOL6 NM_030641.3 F: CCCTGCCAGACCAGGGGACC R: GGAGCGTCATCCTCATCCCTTTGC BCL2A1 NM_001114735.1; F: CCCCGGATGTGGATACCTA NM_004049.3 R: CTAGAAAAGTCATCCAGCCAGA BIRC3 NM_001165.3; F: CCGTCAAGTTCAAGCCAGTTACCC NM_182962.1 R: AGCCCATTTCCACGGCAGCA CCL2 NM_002982.3 F: GCTCGCTCAGCCAGATGCAA R: TCCTGAACCCACTTCTGCTTG CCL4 NM_002984.3 F: CTTCCTCGCAACTTTGTGGT R: GCTTGCTTCTTTTGGTTTGG CCL5 NM_002985.2 F: TGCCTACATTGCCCGCCCAC R: GGGTTGGCACACACTTGGCG CCL20 NM_001130046.1; F: TGACATCAATGCTATCATCTTTCACA NM_004591.2 R: TTTGCGCACACAGACAACTTTT CFB NM_001710.5 F: ATGCCACATACCCCAAAATTTGGGT R: GTTAGTCCCTGACTTCAACTTGTGGT CMPK2 NM_207315.2 F: GCCGGGGCATGGAGAAGACC R: TGGAGGGGCTGGCATCAACCA CSF 2 NM_000758.2 F: CCATGATGGCCAGCCACTACAAGC R: ACTGGCTCCCAGCAGTCAAAGG CSF3 NM_000759.2; F: AAGCTGTGCCACCCCGAGGA NM_172219.1; R: GTGGGACCCAACTCGGGGGA NM_172220.1 CXCL1 NM_001511.2 F: TCAATCCTGCATCCCCCA

216

R: CATAGAATCTTCAAAACTAATGAATAAAT CXCL2 NM_002089.3 F: CGCATCGCCCATGGTTA R: TAGAATCTTCTAAAACAAACAAATAAATA CXCL3 NM_002090.2 F: TCATCGAAAAGATACTGAACAAGGG R: GAAGTGTCAATGATACGCTGATAAGC CXCL5 NM_002994.4 F: CCCAAAATGATCAGTAATCTGCAA R: CAAATTTCCTTCCCGTTCTTCA CXCL10 NM_139089.1 F: TTCCTGCAAGCCAATTTTGTC R: TCTTCTCACCCTTCTTTTTCATTGT EFNA1 NM_004428.2; F: TCACAGTCCTCAGGCCCATGACA NM_182685.1 R: GTGGGGCAGCACTGTGACCG FAM129A NM_052966.2 F: GCTGGACGAGGGCAAGTGCG R: AGGCGCCAATGGTGGCTTGG GAPDH NM_002046 F: TTCACCACCATGGAGAAGGC R: AGGAGGCATTGCTGATGATCT GOS2 NM_015714.3 F: GCGCCGTGCCACTAAGGTCA R: CACGCTGCCCAGCACGTACA ICAM-1 NM_000201.2 F: TGCCCTGATGGGCAGTCAACA R: GCAGCGTAGGGTAAGGTTCTTG IFIT1 NM_001548.3 F: AACCCTGCAGAACGGCTG R: TGTAAAGTGACATCTCAATTGCTCC IFIT3iso1 NM_001549.4 F: GCGTGCCCTACTCTCCCACC R: AGCTGTGGAAGGATTTTCTCCAGGG IFIT3iso2 NM_001031683.2 F: TCAGAACTGCAGGGAAACAGCCA R: AGCTGTGGAAGGATTTTCTCCAGGG IFNGR1 NM_000416.2 F: GAATTTGCTGTATGCCGAGATG R: TGATTTGCTTCTCCTCCTTTCTG IKB /NFKBIA NC_000014.9 F: TGGTGTCCTTGGGTGCTGAT R: GGCAGTCCGGCCATTACA IL1B NM_000576.2 F: TGGCAGAAGTACCTGAGCTCGC R: GCCGCCATCCAGAGGGCAGA IL6 NM_000600.3 F: CCTGAGAAAGGAGACATGTAACAAGA R: GGAAGGTTCAGGTTGTTTTCTGC IL8 NM_000584.2 F: GCAGCTCTGTGTGAAGGTGC R: AAAGGTTTGGAGTATGTCTTTATGCA IL12p35 AF180562.1 F: GGGCCGTCAGCAACATG R: CTTCAGAAGTGCAAGGGTAAAATTC IL32 NM_001012631.1; NM_004221.4;

217

NM_001012718.1; NM_001012636.1; F: GCAGCACCCAGAGCTCACTCC NM_001012635.1; R: AGGCTCCTCGGTTGCGGGAT NM_001012634.1; NM_001012632.1; NM_001012633.1 IRF1 NM_002198.2 Cytoplasmic IRF1: F: CTCACTGCAGCCCCTGCGTC R: TGGGCATGTTGGCTCTGCTGC Unspliced nuclear IRF1: F: GCGACCGCCGAATCG R: TTCGACCCCCCACTTCCT ISG20 NM_002201.4 F: TCCCTGCGGGTGCTGAGTGA R: GCTCCATCGTTGCCCTCGCA LAMB3 NM_000228.2; F: CAGCCAGGCTCCCCAACGTG NM_001127641.1; R: GGCTCGGCTCCTGGCTTCCT NM_001017402.1 MAP3K8 NM_005204.3 F: CGAAGAAAAGAATGGCGTGTAA NM_001244134.1 R: AGCCTGGATTTCCACATCAGA MX1 NM_001144925.1; F: GGCAGCGGGATCGTGACCAG NM_002462.3 R: CCTTCCCCGGCGATGGCATT NFKB2 NM_001077493.1; F: AACCCAAGGAGCCAGCCCCA NM_002502.3; R: CAGCCATATCGAAATCGGAAG NM_001077494.1 NFKBIZ NM_001005474.1; F: GGCTTCTGGCCAAGCTGTGGAT NM_031419.2 R: TCCCCGGGCGTTGGTGTTTG OLR1 NM_002543.3 F: TGGTGCTGGGCATGCAATTATCCC R: GCCGGGCTGAGATCTGTCCCT PI3 NM_002638.3 F: AGCCTGGCTCCTGCCCCATT R: GCAAGGACCGGCTCCCTCTCA PRIC285 NM_001037335.2; F: ACGGTCATTCAGGGCCCACCA NM_033405.3 R: GGCCTCAGCCTGCTCACTGT PTGS2 NM_000963.2 F: GCTGGGCCATGGGGTGGACT R: CCTGCCCCACAGCAAACCGT SOD2 NM_000636.2; F: CGTGGCTGTGGTGGCTTCGG NM_001024465.1; R: CCTGCTGGTGCCGCACACT NM_001024466.1 TFF1 NM_003225.2 F: TCTGCGCCCTGGTCCTGGTG R: GCACACTGGGAGGGCGTGAC TLR2 NM_003264.3 F: GCTGCTCGGCGTTCTCTC R: AAGCAGTGAAAGAGCAATGGG

218

TNF NM_000594.2 Cytoplasmic TNF: F: GTGATCGGCCCCCAGAGGGAA R: TGGAGCTGCCCCTCAGCTTGA Unspliced nuclear TNF: F: TCTCGAACCCCGAGTGACA R: CATCAGCCGGGCTTCAAT TNFAIP3 NM_006290.2 F: AGGCGCTGTTCAGCACGCTC R: CGGGCCATGGGTGTGTCTGT UBD NM_006398.3 F: GGTTTCTGGCCCCTTGTCTGCAG R: ACGCTGTCATATGGGTTGGCATCA U6 NR_004394.1 F: AATTGGAACGATACAGAGAAGATTAGC R: GGAACGCTTCACGAATTTGC ZFP36 NM_003407 F: GCGGGAGTTTTTGCACCA R: GACCGGGCAGTCACTTTGTC

219

Table A4. Primers used for ChIP PCR analysis

Forward (F) and reverse (R) primer sequences (5’-3’) are shown in addition to the accession number for each gene. Primer location for each gene was picked based on the presence of IRF1 binding site as shown in the custom tracks of UCSC genome browser database. All primers were designed using primer designing software (Integrated DNA technologies) and were synthesised by the DNA synthesis lab at the University of Calgary.

Official Human Genome Organization (HUGO) gene nomenclature committee gene symbols have been used for all genes and gene products.

Gene Accession Primer Sequences Symbol Number CMPK2 NM_207315.2 F: GCCAGCGGGAAACGAAAG R: AGGAGGAGAGGCCGGAA CXCL10 NC_000004.12 F: ATTTCCCTCTGCTCCTCTTT R: GAATGGATTGCAACCTTTGTTT IFIT1 NM_001548.3 F: CACCATTGGCTGCTGTTTAG R: CTCCTCTGAGATCTGGCTATTC IFIT3 NC_000010.11 F: GTGGAAACCTCTTCAGCATTTG R: CAGAGAAGCAGGGACTATTTACC hMYOD1(A) NC_000011.10 F: TGCAGGAGATGAAATACTAAGCAAGTA R:AGATTGGAAACTGAGGACTTTAGTTAGAG hOLIG3 NC_000006.12 F: GGCAAGGACAGAGACAATCATA R: CTCTGTGTTCTCGCTTTGGA hMYOG NC_000001.11 F: CCAATGAGACTGAGTGGGTTTTC R: TCACCAGAGAAGACTGCTTTGC

220

Appendix B: Copyright Permission

221