Quick viewing(Text Mode)

Roles of Actin Remodeling Proteins, Gelsolin and Flightless-I in Epidermal Wound Healing

Roles of Actin Remodeling Proteins, Gelsolin and Flightless-I in Epidermal Wound Healing

ROLES OF REMODELING , AND FLIGHTLESS-I IN EPIDERMAL

Huater Chan

Bachelor of Biotechnology (Honours)

Women's and Children's Health Research Institute

And Discipline of

School of Molecular and Biomedical Science

Faculty of Sciences

University of Adelaide

A Thesis submitted for the degree of

Doctor of Philosophy

April 2011 TABLE OF CONTENTS

DECLARATION ...... I

ACKNOWLEDGEMENTS ...... II

ABBREVIATIONS ...... IV

CHAPTER ONE General Introduction

1.1 BACKGROUND...... 1

1.2 WOUND HEALING ...... 2

1.2.1 The process of wound healing ...... 2 1.2.1.1 Inflammatory phase ...... 6 1.2.1.2 Proliferative phase ...... 7 1.2.1.3 Remodeling phase ...... 10

1.3 THE ...... 11

1.3.1 Types of Cytoskeleton ...... 11 1.3.1.1 ...... 13 1.3.1.2 Intermediate filaments ...... 14 1.3.1.3 (Actin) ...... 14 1.3.1.4 Stress fibers ...... 15 1.3.2 Actin dynamics in wound healing ...... 16

1.4 GELSOLIN FAMILY OF PROTEINS ...... 20

1.4.1 Gelsolin ...... 22 1.4.1.1 Gelsolin in Clinical Settings ...... 23 1.4.1.2 Involvement of Gelsolin in Cellular ...... 24 1.4.1.3 Biological Functions of Gelsolin ...... 25 1.4.1.4 Gelsolin and the sex steroid ...... 26 1.4.1.5 Nuclear receptor signaling ...... 28

1.4.2 Flightless-I ...... 29 1.4.2.1 Biological Functions of Flii ...... 30 1.4.2.2 Interactions are mediated through the Flii LRR domain ...... 32 1.4.2.3 Involvement of Flii in cellular processes ...... 33 1.4.2.4 In Vitro and In Vivo Expression of Flii ...... 35 1.4.2.5 Flii protein is a Negative Regulator of Wound Healing ...... 36

1.5 TRANSFORMING GROWTH FACTOR BETA (TGFΒ) ...... 38

1.5.1 TGFβ superfamily ...... 38 1.5.1.1 TGFβ Activation and Structure ...... 40 1.5.1.2 TGFβ Receptor Signaling ...... 44 1.5.1.3 Smad Effector Signaling ...... 45 1.5.1.4 TGFβs regulation in wound healing ...... 48

1.6 MITOGEN-ACTIVATED PROTEIN ...... 50

1.6.1 The MAPK Signaling Pathways ...... 50 1.6.1.1 ERK Pathway ...... 53 1.6.1.2 p38Pathway ...... 54 1.6.1.3 JNK/SAPK Pathway ...... 54 1.6.1.4 Phosphoinositide 3-kinase Signaling pathway ...... 54

1.7 INTEGRATION OF MAPK AND PI3K/AKT WITH TGFΒ SIGNALING PATHWAYS ...... 56

1.8 RESEARCH AIM AND RATIONALE ...... 57

CHAPTER TWO Role of Gelsolin in Androgen Mediated Wound Healing

2.1 INTRODUCTION ...... 60

2.1.1 Implications for the gelsolin superfamily in nuclear receptor signaling ...... 60 2.1.2 Gelsolin and AR Nuclear Signaling ...... 61

2.2 MATERIALS AND METHODS ...... 63

2.2.1 Cells, Culture ...... 63

2.2.2 Scratch Assay ...... 63 2.2.3 Protein Extraction ...... 64 2.2.4 Western Blotting ...... 65 2.2.5 Immunocytochemistry ...... 66 2.2.6 siRNA knockdown Assay ...... 68 2.2.7 RNA Extraction ...... 69 2.2.8 DNase Treatment and RNA Quantitation ...... 70 2.2.9 Complementary Deoxyribonucleic Acid (cDNA) Synthesis ...... 70 2.2.10 Real-Time quantitative-Polymerase Chain Reaction (RTq-PCR) ...... 71 2.2.11 Proliferation Assay ...... 71 2.2.12 Wound Closure Assay ...... 72 2.2.13 Statistical Analysis ...... 73

2.3 RESULTS ...... 74

2.3.1 Gelsolin and AR responses to wounding in vitro ...... 74 2.3.2 Effects of DHT addition in gelsolin and AR localization ...... 80 2.3.3 Time-course localization study of gelsolin and AR in the presence of DHT...... 83 2.3.4 Gelsolin knockdown using silent RNA technology ...... 91 2.3.5 Response to DHT treatment in gelsolin-ablated HFFs ...... 93

2.4 DISCUSSION ...... 98

CHAPTER THREE Flii Deficiency Improves Wound Healing and is Associated with

Higher Expression Ratio of TGFβ3 to TGFβ1

3.1 INTRODUCTION ...... 103

3.1.1 Flii, a negative regulator of wound healing ...... 103 3.1.2 TGFβ and the wound healing process ...... 104

3.2 MATERIALS AND METHODS ...... 106

3.2.1 Flii Heterozygous and Over-expressing Mice Generation ...... 106 3.2.2 Flii+/- mice generation ...... 106 3.2.3 FliiTg/+ mice generation ...... 109

3.2.4 Murine Incisional Wound ...... 111 3.2.5 Histological Processing ...... 112 3.2.6 Immunohistochemistry ...... 113 3.2.7 Cell culture ...... 115 3.2.8 siRNA knockdown ...... 116 3.2.9 RNA extraction ...... 116 3.2.10 DNase Treatment and RNA Quantitation ...... 117 3.2.11 Complementary Deoxyribonucleic Acid (cDNA) Synthesis ...... 118 3.2.12 Real-Time quantitative-Polymerase Chain Reaction (RTq-PCR) ...... 119 3.2.13 Statistical Analysis ...... 119

3.3 RESULTS ...... 121

3.3.1 Time-course analyses of TGFβ1 expression in wounded mice skin ...... 121 3.3.2 Time-course analyses of TGFβ2 expression in wounded mice skin ...... 124 3.3.3 Time-course analyses of TGFβ3 expression in wounded mice skin ...... 129 3.3.4 Time-course analyses of TβRI and TβRII in wounded mice skin ...... 134 3.3.5 Flii knockdown alters TGFβ ...... 139

3.4 DISCUSSION ...... 142

CHAPTER FOUR Flii Modulation of TGFβ Expression is Achieved by Direct Association with

Proteins Involved in the Expression and Activity of TGFβ

4.1 INTRODUCTION ...... 147

4.1.1 Binding properties of Flii protein ...... 147 4.1.2 Properties of the LRR motif ...... 148 4.1.3 The relationship between Flii and TGFβ ...... 149

4.2 MATERIALS AND METHODS ...... 151

4.2.1 Cells, Cell Culture ...... 151 4.2.2 Antibodies ...... 151

4.2.3 Immunocytochemistry ...... 152 4.2.4 Nuclear and cytoplasmic fractionation ...... 153 4.2.5 Immunoprecipitation ...... 154 4.2.6 Western Blotting ...... 155

4.3 RESULTS ...... 157

4.3.1 Flii associates with c-Fos and c-Jun ...... 157 4.3.2 Interactions of Flii with TGFβs ...... 160 4.3.3 Flii interacts with nuclear Akt in wounded cells ...... 169 4.3.4 Flii associates with Smad proteins ...... 172

4.4 DISCUSSION ...... 180

CHAPTER FIVE Regulation of TGFβ by Flii Association with the MAPK Signaling Pathway

5.1 INTRODUCTION ...... 187

5.1.1 Interaction of Flii with Ras ...... 187

5.2 MATERIALS AND METHODS ...... 189

5.2.1 siRNA knockdown ...... 189 5.2.2 RNA extraction ...... 189 5.2.3 DNase Treatment and RNA Quantitation ...... 190 5.2.4 Complementary Deoxyribonucleic Acid (cDNA) Synthesis ...... 191 5.2.5 Real-Time quantitative-Polymerase Chain Reaction (RTq-PCR) ...... 191 5.2.6 Primary Extraction and Culture ...... 192 5.2.7 Immunocytochemistry ...... 194 5.2.8 Secretion Assay ...... 195 5.2.9 Proliferation Assay ...... 195 5.2.10 Fibroblast Outgrowth Assay ...... 196 5.2.11 Statistical Analysis ...... 197

5.3 RESULTS ...... 198

5.3.1 Flii gene knockdown affects expression of AP-1 and Smad proteins ...... 198

5.3.2 TGFβ1 increases collagen secretion in WT, Flii+/- and FliiTg/+ primary ...... 201 5.3.3 TGFβ1 decreases FliiTg/+ fibroblast proliferation ...... 204 5.3.4 Proliferation of primary fibroblasts treated with MAPK inhibitors ...... 207 5.3.5 TGFβ1 decreases migration in Flii+/- fibroblasts ...... 209 5.3.6 MAPK inhibitors decrease fibroblasts outgrowth ...... 214 5.3.7 Comparison of WT, Flii+/- and FliiTg/+ outgrowths treated with MAPK inhibitors ...... 223

5.4 DISCUSSION ...... 229

CHAPTER SIX General Discussion

6.1 DISCUSSION ...... 238

6.1.1 Gelsolin and the AR ...... 238 6.1.2 Flii and TGFβs ...... 240 6.1.3 Future Directions ...... 244

BLIBLIOGRAPHY ...... 238

DECLARATION

I declare that this thesis does not contain without acknowledgement any work submitted previously for any academic award and that to the best of my knowledge and belief it does not contain any material previously published or written by another person except where otherwise acknowledged.

Huater Chan

May 2010

i

ACKNOWLEDGEMENTS

I would like to thank my supervisors, Associate Professor Allison Cowin and Associate

Professor Barry Powell for their invaluable advice and support. I deeply appreciate

their patience, encouragement and support over the years. James Waters, Damien

Adams, Xanthe Strudwick, Zlatko Kopecki, Nadira Ruzehaji and Tony Lin of Wound

Healing Laboratory deserve a big thank you for all the assistance and support they have

provided. Special thanks to our collaborators at the Australian National University at

Canberra, Hugh Campbell, Ruth Arkell, Nicole Thomsen and Jane Hooper-Jones for their support and generosity in providing our laboratory and my project with transgenic mice.

I also would like to thank all the staff and students whom I have the honour to meet at the Women's and Child's Health Research Institute. I am particularly grateful to the past and present students in the PhD room for all their support and friendship throughout the years. I am especially grateful to all the intelligent conversations I had over the years. All the mutual support we gave each other will be solely missed.

ii

Lastly, I would like to thank my friends and family for all their support. Thank you to all my friends for your encouragement and patience over the years. Here I would like to express my deepest appreciation and the biggest thank you to my parents, Hing Chan and Boon Hwa Tan, sister, Yee Send Chan and brother Huasheng Chan. I am eternally grateful and honored to be a part of this wonderful family. I wouldn't be what I am today and could not have completed my studies without them. A special thank you to

Yoke Kim Lee who has accompanied me through the years. I am thankful for her care and concern over the years.

iii

ABBREVIATIONS

A Absorbance

α Alpha

ABP Actin binding protein

APS Ammonium persulfate

AR

β- Beta-tubulin bp

β Beta

BCA Bicinchoninic acid

BSA Bovine serum albumin

°C Degrees celcius

Ca2+ Calcium

CaCl2 Calcium chloride

CaMK-II Calcium/-dependent protein kinase type II

iv

cDNA Complementary deoxyribonucleic acid

CISK Cytokine-independent survival kinase

CO2 Carbon dioxide

Cy3 Cyanine 3 dATP Deoxyadenocine triphosphate dCTP Deoxycytidine triphosphate dGTP Deoxyguanosine triphosphate dTTP Deoxythyamine triphosphate

D0 Day 0 post-wounding

D3 Day 3 post-wounding

D7 Day 7 post-wounding

D14 Day 14 post-wounding

D21 Day 21 post-wounding

DEPC Diethylpyrocarbonate

DHT Dihydrotestosterone

v

DMEM Dulbecco’s modified Eagle’s media

DNA Deoxyribonucleic acid

ECL Enhanced chemical luminescence

ECM

EDTA Ethyldiaminetetraacetic acid

EGF Epidermal growth factor

ELISA linked immunosorbent assay

FCS Fetal calf serum

FITC Fluorescein isothiocyanate

FGF Fibroblast growth factor

FLAP Flightless-I associated protein

Flii Flightless-I protein

FliiTg/+ Flightless-I transgenic

Flii+/- Flightless-I heterozygous knockout

x g Times the force of gravity

vi

g Grams

GTP

H&E Hematoxylin and eosin

HFFs Human foreskin fibroblasts hr Hour

HRP Horse radish peroxidase

H2O2 Hydrogen peroxide

IgG Immunoglobulin-G

IL Interleukin

KCl Potassium chloride

kDa Kilo Daltons

KD Knock down

L Litre

LAP Latency associated peptide

λex Lambda (wavelength) of excitation

vii

λem Lambda (wavelength) of emmitance

LRR Leucine-rich repeats

LTBP latent TGFβ binding protein

M Molar

MAPK Mitogen-activated protein kinase

MgCl2 Magnesium chloride

min Minutes

mM Millimolar

MMP Matrix metalloproteinases

mm2 Millimeter square mRNA Messenger ribonucleic acid

NaCl Sodium chloride

NHS Normal horse serum

NLS Nuclear localization signal nM Nano molar

viii

PBS Phosphate buffered saline

PDGF derived growth factor

PIP2 Phosphatidylinositol 4,5-bisphosphate

PI3K Phosphoinositide 3-kinase

RNA Ribonucleic acid

Rpm Rounds per minute

RTq-PCR Real time quantitative polymerase chain reaction

SDS Sodium dodecylsulphate

SDS-PAGE Sodium dodecylsulphate polyacrylamde gel electrophoresis

SEM Standard error of mean

sec Seconds

siRNA Short interfering ribonucleic acid

TAE Tri(hydroxymethyl)methylamine-acetate-ethylediaminetetraacetic

acid

TβRI Transforming growth factor beta receptor one

ix

TβRII Transforming growth factor beta receptor two

TEMED N,N,N,N-tetramethylethylenediamine

TGFβ1 Transforming growth factor beta one

TGFβ2 Transforming growth factor beta two

TGFβ3 Transforming growth factor beta three

TNF Tumour factor

Tris Tri(hydroxymethyl)methylamine

TRS Target retrieval solution

µg Microgram

µl Microlitre

µm Micrometer

µM Micromolar

VEGF Vascular endothelial growth factor

WT Wild-type

x

WST-1 2-(4-iodophenyl)-3-(4-nitropheyl)-5-(2,4-disulfophenyl)-2H-

tetrazolium

x Times

x g Times the force of gravity

% Percent

= Equals

+ Plus

xi

CHAPTER ONE

General Introduction

1.1 Background

Scar contractures arising from abnormal wound healing continue to pose a significant worldwide clinical problem. Elective operations and operations due to trauma in the developed world account for more than 100 million patients acquiring each year, which can lead to significant health problems. In addition, burn is the most common household injury and results in more than 5 million patients worldwide acquiring burn scars each year, 70% of which occur in children. The problems are not just limited to physical and psychological trauma of the individual but also extend outwards and place a significant burden on families and communities for treatments.

Scarring results in loss of movements, restricted growth and deformity and is particularly important for children whose development often requires regular surgical correction. The long term economic impact involves loss of wages, constant health care cost and loss of skills as a result from scarring deformities. Unfortunately, current wound healing treatments such as garments, dressings or steroidal injections are sub- optimal. As a result, more research into cell based therapies or pharmaceutical drugs to improve wound healing must be conducted.

1

1.2 WOUND HEALING

1.2.1 The process of wound healing

Wound healing is a restoration process in which the skin repairs itself after injury. In normal skin, the epidermis and dermis act as a protective barrier against the external environment and in the event that this barrier is breached, the systematic physiological process begins to re-establish the integrity of the protective barrier. Molecularly, cutaneous wound healing is a highly complex process involving biochemical reactions and interactions amongst cells and cytokines leading to the restoration, integrity and function of the skin. Cellular migration is one of the most important aspects of wound healing to restore the protective barrier. This process involves the actin cytoskeletal proteins which are regulated by the gelsolin family of actin binding proteins such as gelsolin and Flightless-I, which will be discussed later in this chapter.

The coordination of multiple cellular processes is a very delicate process, which ultimately determines the resultant formation from an unnoticeable hairline graze to keloid or hypertrophic scar contractures (Figure 1.1) at the opposite end of the spectrum. The process of wound healing can be divided into 3 distinct but interrelated phases (Figure 1.2), inflammatory phase, proliferative phase and remodeling phase 1.

2

A C

B D

Figure 1.1

Typical outcomes of abnormal wound healing processes in which repair is non- optimal. (A) Hypertrophic scars are usually raised and limited to the margins of the original wound. (B) Keloid scars infiltrate and invade the surrounding normal tissue area. Histological sections of hypertrophic (C) and keloid (D) scars. Type III collagen bundles are flatter but are predominantly parallel to the epithelial surface in hypertrophic scars. α-SMA (alpha - Smooth Muscle Actin) positive myofibroblasts are

present in hypertrophic scars. However, early collagen fibers in keloid scars are

arranged haphazardly to the epithelial surface possibly due to the absence of α-SMA

positive myofibroblasts.

Adapted from Santucci et al, 2001& Wolfram et al, 2009 6,7

3

NOTE: This figure is included on page 4 of the print copy of the thesis held in the University of Adelaide Library.

Adapted from Gurtner et al, 2008 8

4

Figure 1.2

Classic stages of the wound healing process. There are three phases of adult wound healing; (A), proliferation (B) and remodeling (C). The inflammatory phase, hemostasis, is restored by the formation of fibrin clots and, during this time, potent cytokines and chemokines are released to attract inflammatory cells into the wound site to prevent infection. The proliferative phase occurs about 2 – 10 days post- injury. It begins when fibroblasts migrate into the wound space to proliferate and differentiate into collagen producing myofibroblasts. The production of new ECM

(Extra Cellular Matrix) proteins allows angiogenesis and the subsequent migration of epithelial cells from nearby unwounded tissue into the wound site to re-epithelialize.

The remodeling phase involves the re-organization of the tissue matrix and lasts approximately for a year post-wounding. The contracted and re-epithelialized wound is slightly raised and damaged skin appendages are not restored. Dermal laid down by fibroblasts are disorganized and the collagens begin to mature and are re- organized into thicker bundles that inevitably cause the formation of scar tissue.

5

1.2.1.1 Inflammatory phase

Wound healing is initiated immediately in response to injury and the first action taken

is the restoration of hemostasis, characterized by the formation of a clot to prevent

further blood loss, and the onset of inflammation. The formation of a clot not only

provides a first barrier against infection but also forms part of the provisional matrix

which is required for incoming inflammatory cells into the wound site. Immediately

upon injury, fibrinogen is first cleaved by thrombin to produce fibrin. Fibrin monomers

then cross-link with one another to form a scaffold which in turn bind directly to

to form a clot 2. The formation of fibrin is integral in the early phase of wound

healing as it binds to different cell types such as monocytes, and fibroblasts

via cell surface integrin receptors. Fibrin can also interact with several growth factors

and cytokines such as fibroblast growth factors (FGF) and insulin-like growth factors

(IGF), which affects cellular migration and proliferation, as well as extracellular matrix

(ECM) production 3. Concurrently with the formation of a clot, platelets are also

activated by thrombin upon injury to release granules which contain a range of growth

factors vital in wound healing such as platelet-derived growth factors (PDGF), transforming growth factors (TGF) and epidermal growth factors (EGF) 4.

In addition, platelets also stimulate vasodilation and increase permeability to allow inflammatory cells to enter the wound site 5. This process is regulated by numerous

6

cytokines and growth factors such as mast cells that produce histamines and

complement factors that assist in vasodilation and act as chemoattractants 9. As a result, vasodilation and increased permeability cause a cascade of inflammatory cell influx and differentiation to supplement the release of more cytokines and growth factors at the wound site 5. Neutrophils are the first population of inflammatory cells that enter the wound site to clear up cell debris and invading bacteria 10. Neutrophils

also produce several proinflammatory cytokines such as interleukins (IL-1α, IL-1β, IL-6) and tumour necrosis factor – α (TNFα) 11. These factors attract circulating monocytes which migrate into the wound site where they are activated and subsequently differentiate into , where the process of continues in conjunction with the production of chemoattractants, , elastin, complement factors and TGFβ 12. These factors together recruit more macrophages and fibroblasts to

the wound site. One of the most important cytokines released by the macrophages is

TGFβ which stimulates dermal fibroblasts to differentiate into myofibroblasts. The myofibroblasts then produce collagen which forms part of the provisional extracellular matrix 3. The migration of fibroblasts into the wound space marks the beginning of the

proliferative phase in wound healing.

1.2.1.2 Proliferative phase

Fibroblasts are important mesenchymal cells which play a dual role in the wound

7

healing process 10. Firstly, fibroblasts produce collagen based ECM that eventually replaces the fibrin matrix, therefore regulating both growth and function of other cell types. Secondly, differentiated fibroblasts are also involved in wound contraction. The migration of fibroblasts into the wound site occurs in a well coordinated manner by navigation along the provisional matrix fibers and not in a haphazard manner 13. It is

therefore important to note that the arrangement of ECM in wound healing is crucial, as

many complications such as scarring arise out of disorganized ECM arrangement. The

interactions between the fibroblasts and the ECM are mediated by cell surface integrins

that bind to several matrix components such as fibrin and fibronectin 14. IL-1 and TNF-α

production by macrophages also induces the production of matrix metalloproteinases

(MMP) that clear away damaged ECM by removing the inflammatory debris, which

then allows the migration of proliferative cells into the wound site 15. Initiation of

permanent ECM production begins at the arrival of fibroblasts into the wound space.

The new ECM is comprised of mainly collagen which is predominantly synthesized in

fibroblasts as procollagen and exported out of the cell 16. In humans, collagen

production typically starts at day two post-injury and peaks between day five and

seven. The production of collagen matrix and the continuation of fibroblast migration

results in the development of which provides a template for tissue

growth and formation of myofibroblastic cells 17. A range of growth factors and

8

cytokines including PDGF, EGF, FGF TNF-α, interferon-γ and most importantly TGFβ

are involved in the differentiation of fibroblasts into myofibrobasts which are major

contributors in extracellular matrix deposition. These specialized fibroblasts acquire

properties characteristic of smooth muscle cells and have the ability to generate

mechanical forces which result in lamellipodia formation and subsequent wound

contraction.

The formation of ECM is the first criteria that must be fulfilled before angiogenesis can

occur, as the ECM provides structural support in addition to a repository of important

growth factors for invading 18. Angiogenesis is the process where new blood

vessels are formed from neighboring intact capillaries. The movement of new capillaries

into the wound site is similar to the migration of fibroblasts that uses the established

ECM as guidance. Multiple growth factors are involved in this entire process including

VEGF, FGF, angiopoietin and TGFβ.

Re-epithelialization involves the migration of epithelial cells from the wound edge using components from the ECM such as collagen and fibronectin as structural foundations which ultimately results in re-establishment of intact epidermis over the newly formed granulation tissue. At the leading edge of the migrating tip are the keratinocytes. As basal keratinocytes migrate over the granulation tissue, they leave a

9

stratified layer of proliferating keratinocytes and the process continues until

keratinocytes from opposite ends come in contact 1. Rapid re-epithelialization is desirable as it leads to the restoration of the skin’s function as a barrier to defend against micro-organisms, protect from trauma and prevent water loss.

1.2.1.3 Remodeling phase

The remodeling phase involves re-organization of the tissue matrix by means of removal of fibronectin and hyaluronic acid and the replacement of a stronger and more

organized ECM framework composed of collagens 19. Degradation of the initial matrix

and regulation of the new collagen matrix involves the family of matrix

metalloproteinases (MMP), which require controlled expression of exact combinations

of MMPs regulated in part by PDGF, IL-1 and TGFβ 20. Regulation of collagen requires a

balance between collagen production, collagen breakdown and collagen remodeling.

Over time, the proportion of collagen I content in the granulation tissue will increase

with a corresponding decrease of collagen III until it returns to the basal levels of

collagen I to collagen III ratio of 90% : 10% in unwounded skin 19.

Maturation of collagen fibers occurs approximately six to ten days after injury where

the deposited collagens are absorbed and re-organized into thicker bundles parallel to

the skin, correlating with increased tensile strength 12. Studies suggest that new collagen

10

fibres are orientated in an organized manner which is characteristic of the remodeling

phase of wound healing and results in scar formation, ranging from fine lines to

widespread scars 21. Scar formation is the inevitable outcome in the final remodeling

phase of wound healing in adult humans. One of the causes is the influence of fibrotic

agents, causing fibrosis to take place during the fibro-proliferative response. The main

cause of scarring is the continual presence of fibroblasts that over synthesize collagen

leading to hypertrophic scar morphology 12.

1.3 The Cytoskeleton

1.3.1 Types of Cytoskeleton

The cytoskeleton is a complex and highly dynamic network of microfibers found in the

cell’s . The cytoskeleton was originally thought to provide cellular mechanical

strength, locomotion, support organelles and maintain cell shape. However, it is now well-known that the cytoskeleton is also involved in other processes as well as pathologies arising from cytoskeletal defects 22. The structural integrity of eukaryotic

cells are made up of three molecular building blocks, microtubules, microfilaments and

intermediate filaments (Figure 1.3) differentiated by their sizes from large to small

respectively 23. Each of these cytoskeletal proteins has their own functional roles but

together they function synergistically with each other to withstand higher mechanical

11

Cell membrane

Nucleus

Centrosome

Microtubule Intermediate filaments Actin

Figure 1.3

Illustrative figure and immunofluorescent staining of microtubules, intermediate

filaments and actin within a typical cell.

Microtubule image obtained from Monton et al, 2009 24

Microtubules 12

stress25,26. The cytoskeletal properties in epithelial cells are especially important in

wound healing as they not only sense the environment but also allow migration from

healthy unwounded tissue into the wound site to re-establish epithelial integrity.

1.3.1.1 Microtubules

Microtubules are large fibrous cytoskeletal structures measuring approximately 25nm in diameter. They are arranged in straight, hollow, cylindrical aggregates made up of alpha and beta tubulin dimers assembled together 27. Microtubules are highly dynamic

and participate in a wide range of cell activities, mostly involving motion but also

including cellular division, morphogenesis and organelle transport. Being an important molecule in cellular motility, the highest concentration of microtubules are found at locomotive structures such as the cilia and flagella of motile cells 27. Cellular

propagation is achieved by the continuous of tubulin dimers at the plus

end and depolymerization at the minus end of a polarised cell. This process utilizes

energy by means of GTP hydrolysis which is expected since tubulin has GTPase

activity. The movement of cellular components such as vesicles takes place along the

microtubule network. This process requires two microtubule motors, , which

moves toward the plus end of the microtubules and dynesin, which moves toward the

minus end of the microtubules 28.

13

1.3.1.2 Intermediate filaments

Intermediate filaments are the second largest cytoskeletal structure after the microtubules, measuring approximately 10nm in diameter 23. Intermediate filaments consist of a conserved central α-helical domain flanked by variable N and C termini that contribute to the diversity of the family 29. The α-helical rod shape arrangement contains seven-residue repeats that interact with other intermediate filament protein to form a coiled-coil dimer 30. This structural configuration allows the intermediate filaments to interact with a range of protein complexes at the cell surface, including desmosomes, hemidesmosomes, focal adhesions and the extracellular matrix

31, 32. The interaction with cell surface molecules suggests that intermediate filaments may act as a sensory mechanism that relays signals from the cell surface to the surface of the nucleus.

1.3.1.3 Microfilaments (Actin)

Microfilaments, also known as actin, are the smallest of the three cytoskeletal proteins and the most important during wound healing. Actin is a highly conserved eukaryotic molecule expressed in animals, plants and fungi. Feuer et al, 1948 33 first identified actin as a major muscle component. Since then, actin has been identified in multiple important cellular processes including cellular motility, cell division and structure.

There are two main isoforms of actin, unpolymerized globular monomeric actin (G-

14

actin) and polymerized, filamentous actin bundle (F-actin); both exist in equal amounts in vivo 34. F-actin assembly is an energy requiring process in the form of ATP hydrolysis

35. Double-stranded, helical F-actin is always in a constant state of flux with G-actin

monomers, being added to the ‘plus’ end and depolymerized at the ‘minus’ end.

Comparison of different actin structures showed that while their overall configurations

are analogous, they show considerable local differences due to the intrinsic dynamics of

actin filaments modulated by different effector molecules and interactions with other

proteins to form a complex 36.

1.3.1.4 Stress fibers

Stress fibres are contractile actomyosin structures composed of mainly unorganized

actin filament bundles. This is the basic cellular structure that provides the contractile

force required for cell morphogenesis and migration 37. The constant contraction of

stress fibres are in balance with cellular adhesion strength as it results in a more stable

formation of actin bundles that maintain a constant length even under tension 38. There

are three classes of stress fibres in mammals. The first class is the ventral stress fibres

that associate with focal adhesions at both ends and are responsible for tail retraction 39.

The second class is the transverse arcs found from the leading edge to the cell centre

and they do not directly associate with focal adhesions but connect to the substrate via

the dorsal stress fibres which are the third class of stress fibres 39,40. Dorsal stress fibres

15

associate themselves from the focal adhesion sites at one end to the dorsal section of the cell at the opposite end and therefore efficiently propagate contractile forces from the cell to the substrate 37. Stress fibres highlight the dynamic capabilities of actin filaments and their importance in cell migration.

1.3.2 Actin dynamics in wound healing

In the event of an injury, surrounding motile cells such as fibroblasts migrate towards each other during wound healing. These wound edge motile cells are induced by various extracellular signals to undergo a transition from a non-polarized to a polarized state 35. Once polarized, cells have the ability to migrate by forming and extending protrusive structures, known as lamellipodia and filopodia. Lamellipodia are large sheets which consist of branched actin filaments at the front of a leading edge cell and filopodia are long, parallel actin filaments that protrude beyond the 41.

Evidently, actin in this situation develops two different organizations for lamellipodial formation and filopodial protusion. In lamellipodia, small GTPases and Rac activate proteins of the Wiskott-Aldrich syndrome family such as Wiskott-Aldrich syndrome protein (WASP) and WASP-family verprolin-homologous protein (WAVE) which in turn activates Actin-related protein (Arp) 2/3 complex 42. Activated Arp2/3 complex will lead to the nucleation of actin filament branches that results in a broad network of actin filaments 43. Cdc42 in filopodia however, promotes parallel linear actin

16

polymerization, aided by , vasodilator-stimulated phosphoprotein (VASP) and

fascin 44. Clearly, the formation of these protrusion structures is tightly regulated and

cellular migration is dependent on the balance between polymerization on the ‘plus’

end and depolymerization at the ‘minus’ end of F-actin. Therefore to facilitate cellular

motility, the process of polymerization and depolymerization has to be coordinated

such that, on average, the filament moves forward in one direction.

Cell-matrix adhesion is another factor determining migration. During migration, cells

form protrusions that require new attachment sites or focal contacts at the leading edge

to provide traction between the actin cytoskeleton and the underlying substrate 45. The

formation of new focal adhesion contacts is mediated by integrins which are a family of

heterodimeric transmembrane receptors that connect the extracellular matrix to the

actin cytoskeleton 46. Focal complexes act as a "clutch handle" to connect the integrins to

the actin cytoskeleton and in the early stages consist of dynamic proteins such as β3-

integrin, and paxillin (Figure 1.4). In addition to providing traction, cell-matrix adhesions also act as extracellular sensors. The force generated by the actomyosin contraction induces focal complex maturation which allows further recruitment of actin binding proteins such as α-, VASP and FAK and further actin polymerization 47-

49. These protein complexes are localized to the leading edge

17

NOTE: This figure is included on page 18 of the print copy of the thesis held in the University of Adelaide Library.

Adapted from Le Clainche et al, 2008 50

18

Figure 1.4

Cellular adhesion acts as a "clutch handle" to connect the integrins to the cytoskeleton.

This allows the conversion of the force generated by actin polymerization into protusion. The actin network is represented in grey and newly formed actin is represented in pink. Adhesion molecules are represented in blue. Cellular substrate is represented by the red lines. The cell boundary is represented by the yellow lines. (A)

No connection of the adhesion molecules and the actin cytoskeleton. In this scenario, no protusion occurs as actin polymerization is converted into retrograde flow. (B)

Connection of the polymerizing actin network to the substrate via the adhesion molecules. In this scenario, the connection between the polymerizing actin network and the substrate provides traction that enables cellular protusion.

19

of cells which assemble and disassemble during cellular migration. These events,

coupled with the forward motion, occur with the simultaneous retraction of attachment

sites at the trailing end of the cell, allowing the cell to pull its rear, therefore driving the

cell in a single direction. This process is known as ‘treadmilling’ which allow cells to

migrate into the wound site (Figure 1.4). ‘Treadmilling’ is a highly complex process that

requires the coordination of a set of proteins including but not limited to actin

depolymerizing factor (ADF) or cofilin, and capping proteins such as the

gelsolin family of actin binding proteins 35.

1.4 Gelsolin family of Actin Remodeling Proteins

Cytoskeletal rearrangement is fundamental in wound healing to facilitate cell

movement and its process involves a wide range of proteins. Among these is the

gelsolin superfamily of actin binding proteins (ABP). The gelsolin superfamily is a

conserved family of proteins present in mammalian cells and is characterized by having either three or six homologous gelsolin-like structural domains known as G1-G6

segmental domains (Figure 1.5A). Early studies of gelsolin, the founding member of the

family, identified three actin binding regions, a calcium independent strong monomer

binding fragment (G1), a calcium independent filament binding fragment (G2-3) and a calcium dependent monomer binding fragment (G4-6) 51.

20

NOTE: This figure is included on page 21 of the print copy of the thesis held in

the University of Adelaide Library.

Figure 1.5

(A) Schematic representation of human domain structure and residue

G1-G6. Ca2+, phosphatidylinositol 4,5-bisphosphate (PIP2) and actin binding segments are shown. (B) Gelsolin superfamily members include , supervillin and flightless-I that have additional unique domains beyond gelsolin G1-G6 domains. CapG which has only 3 gelsolin domains is also a member of the superfamily. ABD, actin binding protein. NLS, nuclear localization signal.

Adapted and modified from Sun et al, 1999 52

21

Other members of the family such as villin and supervillin have additional actin

binding domains that provide a unique function to each member 53. Supervillin has an

N-terminal addition which is capable of protein-protein interactions and nuclear

localization. CapG has undergone evolutionary truncations and has only the G1-G3

gelsolin domains but still retains full actin severing properties 54. The Flightless-I protein

(Flii) contains an N-terminal leucine rich repeat (LRR) region that is capable of protein- protein interactions. Taken together, members of the gelsolin family have roles other than just actin remodeling owing to their specific additional domains.

Previously, it was thought that there was a potential for redundancy due to the homology between members of the gelsolin family. This is true to some extent owing to the conserved gelsolin-like domains. However, research has shown otherwise. For example, expression patterns of gelsolin and CapG are complementary to one another indicating that they have distinct in vivo functions 55. In fact, a CapG and gelsolin double

knockout study demonstrated that CapG is highly essential in actin-based

motility that is distinct from gelsolin and its functions include macrophage receptor mediated ruffling, IgG complement and phagocytosis 56.

1.4.1 Gelsolin

Gelsolin is the founding member of the family and regulates the dynamics of

22

filamentous actin by means of binding, severing and capping actin filaments 52. This

process is initiated by Ca2+ which causes a conformational change that allows gelsolin to

bind to actin and severing it while remaining bound to the barbed end. Uncapping of

actin occurs in the presence of phosphatidylinositol which bind to gelsolin thereby uncapping and exposing the barb ends for further actin polymerization 57.

Other modes of action that are also known to affect activity include phosphorylation

and binding to (ATP) 58,59.

1.4.1.1 Gelsolin in Clinical Settings

Recent research has implicated gelsolin in various diseases including a systemic

disorder known as Hereditary gelsolin 60,61, Meretoja's syndrome (lattice

) 62 and a role in Alzheimer's disease 63-65. Furthermore, over-

expression of gelsolin is linked to a negative prognosis in a subset of breast tumours

which show higher expression of tyrosine receptor kinase erbB-2 and EGFR 66.

Interestingly, excluding the above subset of carcinomas, gelsolin is generally considered

to be a tumour suppressor 67. It is reported that tumour progression is correlated with

decreased expression of gelsolin and that over-expression of gelsolin inhibits carcinogenesis. These observations suggest that gelsolin possesses complex multi-role

properties other than as a downstream actin remodeling molecule.

23

1.4.1.2 Involvement of Gelsolin in Cellular Apoptosis

Apart from being a simple cytoskeletal actin modulator, gelsolin is also involved in

other cellular processes such as being a contributor to cellular apoptosis 68 and the

regulation of Rac proteins 69. Gelsolin is a substrate for caspase-3 which is the core effector caspase activated during apoptosis 70. Gelsolin severing of the actin

cytoskeleton is calcium dependent under normal cellular conditions. However, when

cleaved by caspase-3 to produce the amino-terminal cleavage product (residues 1-352),

gelsolin severing of the actin cytoskeleton becomes calcium independent and results in

the disassembly of the membrane cytoskeleton, which is characteristic of apoptosis 68.

Over-expression of this gelsolin fragment in cells results in apoptosis whereas gelsolin

knockout neutrophils exhibited a delay in the onset of apoptosis more slowly than wild-

type neutrophils 71. These findings highlighted the role of gelsolin in mediating the

apoptotic pathway.

Conversely, a pro-apoptotic role of gelsolin has also been identified. This is supported

by the fact that most cancer cells expressed significantly lower levels of gelsolin 72-74. In

addition, research has shown that over-expression of gelsolin inhibits apoptosis 71. This

is consistent with findings from another group 75 showing the inhibition of apoptosis in

stimulated Jurkat cells (human T-cell line) following gelsolin over-expression at several

times the normal levels. Furthermore, over-expression of a gelsolin mutant form that is

24

more sensitive to phosphoinositide, suppresses Ras transformation 76. Currently, the conflicting role of gelsolin in apoptosis remains to be elucidated. However, one possible explanation for the conflicting observations is that gelsolin may interact with other intracellular proteins such as phosphoinositides, which as a complex, function as a competitive inhibitor of caspase-3 77. These findings which implicate gelsolin in apoptosis and carcinogenesis highlight the complexity in the balance between the multiple regulatory role of gelsolin.

1.4.1.3 Biological Functions of Gelsolin

Genetic knockout of gelsolin display a mild phenotype allowing the mouse to develop into adulthood 78. Regardless of the mild phenotype accompanied with genetic knockouts, the effect of manipulating gelsolin expression is significant at the cellular level. Gelsolin null skin tissue has pronounced actin stress fibers due to the limited ability to sever and remodel actin filaments 78. As a result, the lack of gelsolin causes poor ruffling of actin filaments in response to growth factors 69 as well as the restriction of platelet activation by actin severing, subsequently hindering the clotting mechanism

78. In addition, cellular motility is heavily reliant on actin cytoskeletal rearrangement.

For example, the formation of lamellipodial and filopodial protusions of neuronal growth cones are induced mainly by adseverin and the presence of Ca2+ but in gelsolin null mice the retractions of the protusions are severely delayed, slowing cellular

25

motility 79. Furthermore, the motility of migratory cells such as fibroblasts is also

negatively affected in gelsolin null mice 78. Gelsolin-null macrophages also exhibit

impaired IgG-mediated phagocytosis 80. As a consequence, gelsolin null mice exhibit

defective chemotaxis and slower wound healing, consistent with reduced cellular

motility. These findings indicate the importance of gelsolin in the regulation of actin

dynamics.

1.4.1.4 Gelsolin and the sex steroid hormones

Previous studies have implicated gelsolin in the hormone mediated wound healing

process 53,81,82. It is known that androgens and estrogens are important regulators of

wound repair, highlighted by the improvement of healing in young females versus elderly males and females 83. However, sex steroid hormones are better understood in the development of primary and secondary sexual characteristics. In humans, adrenal

cortex and the primary sexual organs, ovaries and testes, are major production sources

of sex steroid hormone precursors. Both estrogen and androgen are synthesized from

an inactive common steroidal precursor, dehydroepiandrosterone that is secreted in

large amounts by the adrenal cortex 84. The formation of the inactive form allows target

tissues to control the formation and metabolism of the sex steroid hormones according

to their requirements. Activation requires the initial conversion of

ehydroepiandrosterone to androstenedione by 3β-hydroxysteroid dehydrogenase.

26

Several namely, 3β-hydroxysteroid dehydrogenase, 17β-hydroxysteroid

dehydrogenase, 5α-reductase and aromatase subsequently convert androstenedione to

the naturally more potent bioactive estrogen, 17β-estradiol and androgen,

Dihydrotestosterone (DHT) which are involved in downstream biological processes.

Downstream effects of the sex hormones are mediated through the nuclear hormone

receptors which regulate the expression of target . Post-menopausal women suffer

from impaired wound healing and this is correlated with lower physiological levels of

estrogen. This is highlighted by a prolonged inflammatory response, higher levels and reduced matrix deposition 85. Closer studies revealed that topical estrogen is

able to repair skin atrophy, wrinkles and dryness in post-menopausal women and is

associated with accelerated wound repair, highlighted by a reduced inflammatory

response 86 and enhanced extracellular matrix deposition 83. Topical treatment with

estrogen was also found to be beneficial by reversing the reduced healing rate in elderly

women 87. Male counterparts also benefit from this treatment which led to significantly

lower inflammatory responses during wound healing.

In contrast, androgens impede wound healing and are associated with increased

inflammation and wound immune cell dysfunction 88. Furthermore, murine studies

done byAshcroft et al, 2002 89 and Gilliver et al, 2006 90 reported that castration of male

27

mice or treatment with AR antagonists resulted in a marked improvement of wound repair. This is in part due to the decrease in overall inflammation by inhibiting the upregulation of proinflammatory cytokines by endogenous androgens. As a result, target inhibition of 5α-DHT expression could be used therapeutically for the acceleration and improvement of wound healing.

Taken together, the findings show that while estrogen positively regulates the wound healing process, androgens showed detrimental effects. Therefore, being male is considered to be a risk factor for abnormal wound healing particularly in the elderly 91.

Clearly the complexities of sex hormones are not restricted to the sole purpose of regulating and maintaining sexual characteristics, it also significantly affects the process of wound healing. Therefore, it is safe to assume that while the underlying wound healing process in males and females are similar, there are differences in the mechanism due to the variation in hormonal makeup in males and females. The implications for sex hormones in wound healing have only been established recently and the mechanism of their actions remains unclear.

1.4.1.5 Nuclear hormone receptor signaling

Androgens (testosterone or its metabolite DHT) act through binding to the Androgen

Receptor (AR), which is a ligand inducible nuclear receptor. The AR is an 110kDa

28

protein belonging to the nuclear receptor family which upon binding to a ligand, typically DHT, translocates into the nucleus and binds to androgen response elements

(AREs). In addition to its primary role, AR also plays important roles in the sex hormone mediated wound repair process.

Nuclear receptors (NR) constitute a family of transcription factors that is regulated by diverse ligands ranging from steroid hormones, lipids to retinoids. NRs mediate their effects by binding to response elements at their target promoter regions (reviewed in 92.

NRs have a centrally located, highly conserved DNA-binding domain (DBD) characteristic of all NRs. Located in the C-terminus of the hormone binding domain and the N-terminus lies the activator function 1 (AF-1) and activator function 2 (AF-2) that are primarily responsible for transcription activation by hormone-activated bound NR.

Their full activity is dependent on ligand binding and the target gene promoter. The majority of ligands bind NRs as dimers that either regulate transcription directly by binding to enhancer elements at the gene promoter, or indirectly by potentiating

nuclear translocation and DNA binding efficiency.

1.4.2 Flightless-I

Flii is a member of the gelsolin superfamily and has highly conserved homologues

sharing 52% and 69% protein sequence similarity between Caenorhabditis elegans and

29

human respectively 53. The human Flii gene spans 14kb of genomic DNA and contains more introns than human gelsolin or villin 54. The carboxy terminal half of human Flii

has 31% identity and 52% similarity to human gelsolin (Figure 1.5B) 93. The Flii gene is

located in the subdivision 19F on the X- which encodes a 1256 peptide with a molecular weight of 143, 672 Da 94.

1.4.2.1 Biological Functions of Flii

The name Flii was given because point mutations in the gene disrupt the structural

organization of Drosophila melanogastor’s indirect flight muscle myofibrils and result in a

flightless phenotype 95. Drosophila melanogastor eggs lacking maternally supplied Flii

showed incomplete cellularization and impaired gastrulation, associated with a

disorganized actin cytoskeleton (Figure 1.6) 96. The flightlessness and the incomplete

cellularization phenotypes indicated that Flii is required for proper actin organization

during myogenesis and embryogenesis respectively. In humans, the Flii has been

mapped to a region deleted in the Smith-Magenis Syndrome which is associated with

developmental and psychological abnormalities 97.

Homozygous knockout studies of Flii revealed more important roles of Flii during

development. In Drosophila, homozygous knockout of Flii resulted in the disruption of

the syncytial blastoderm during cellularization leading to early embryonic

30

Figure 1.6

Cross-sections showing abnormal cellularization of Flii mutant Drosophila embryos. (a, c, e) Wild-type embryo. (b, d, f) Flii mutant embryo. No significant differences between wild-type and Flii mutant embryos are visible during syncytial blastoderm stages (a-b).

However, when cellularization occurs and the nuclei begin to elongate, the layer of nuclei in Flii mutant embryos begins to appear disorganized (c-d). During gastrulation, nuclei align themselves to face the outside of the embryo (e). In Flii mutant embryos, the nuclei move out of the incompletely cellularized peripheral layer of cytoplasm into the interior of the egg. Bar = 50µm in (b).

Obtained from Straub et al, 1996 96.

31

developmental arrest 98. In murine models, homozygous knockout is embryonic lethal,

with development arresting at a stage preceding gastrulation. Using a human Flii

transgene, normal development of Flii homozygous knockout mouse embryos can be

restored 98. Coupled with the fact that Flii heterozygous mice exhibit no impediments during development, this indicates that a single copy of Flii gene is adequate for normal gene function 98. A homolog of Flii has been implicated in the establishment of the

anterior-posterior polarity in Caenorhabditis elegans 99. It was found that the Flii

homolog also regulates the cytokinesis of somatic cells and the development of

germline. These findings indicated the importance of Flii in performing essential

functions during early embryogenesis in Drosophila, C. elegans and mammals.

1.4.2.2 Protein Interactions are mediated through the Flii LRR domain

The Flii protein represents an interesting fusion of two functionally distinct protein

families’ evolutionary gene insertion events. Flii is a functional filamentous actin

severing protein but, unlike gelsolin, it functions independently of Ca2+ 100. The 5’ end of

Flii protein consists of 16 tandem repeats of 23 amino acid LRR fused to the gelsolin-like

G1-G6 segmental domain by a linker of approximately 100 amino acid residues, making

it a unique member of the gelsolin superfamily (Figure 1.5B). Like many other LRR

containing proteins, the Flii LRR domain may form a hydrophobic, curved solenoid

structure 101 that is particularly suitable for protein-protein as well as protein-

32

interactions involved in , either by binding directly as a ligand or as

a regulator to mediate receptor-ligand binding affinity 102. The first binding partner of

Flii discovered is known as the Flightless-I LRR associated protein (FLAP) of size 626

amino acids 103.

1.4.2.3 Involvement of Flii in cellular processes

As the functions of Flii are still largely unknown, interactions of Flii in cellular

processes continue to be revealed. Recent research has reported that Flii directly and

preferentially interacts with the active form of calcium/calmodulin-dependent protein

kinase type II (CaMK-II), a protein kinase involved in the progression of the cell cycle

104. When CaMK-II was inhibited, Flii over-expression suppressed the transcription of β- dependent transcriptional reporters but the suppression of Flii expression enhanced β-catenin transcription. Furthermore, the mammalian Flii has been identified to be involved in the nuclear receptor signaling by directly associating with estrogen and thyroid hormone receptors as well as their co-activating transcriptional factors,

CARM1 and GRIP1 105. In this case, Flii localizes to the promoter region and functioned as a co-activator protein that actively recruit and coordinate other transcriptional factors

such as p160, mediator and the SWI/SNF complex 106. This allows the assembly of a

larger estrogen receptor associated co-activator complex onto the promoter region of an

estrogen inducible gene. On the other hand, a cytokine-independent survival kinase

33

(CISK) is a downstream phosphoinositol 3-kinase target has been identified to interact with Flii to regulate the estrogen receptor 107. In the study, Flii was found to be a CISK

substrate where it was shown that CISK could phosphorylate Flii at residues Ser436 and

Thr818. This indicated that Flii could be post-translationally modified and is of

importance in determining its multifunctional properties. These findings suggest a role

of Flii in the regulation of the cell cycle.

Flii has also been implicated in the immune response. It was revealed that Flii regulated

pro-inflammatory caspase-1 and capase-11 by direct association and modulation of their intracellular localization and activity 108. The regulation of the pro-inflammatory

caspases impact on the maturation of interleukin-1β 108, therefore potentially affecting

its activity in processes such as wound healing. The role of Flii in the immune response

is also reported by Wang et al (2006). The authors showed that Flii directly interacts

with MyD88 which is an immediate downstream adaptor protein of Toll-like receptors

(TLR) involved in the regulation of signaling specificity in the .

This interaction prevents the formation of TLR4-MyD88 complex which subsequently inhibited the activation of NFkB.

34

1.4.2.4 In Vitro and In Vivo Expression of Flii

Flii can be induced by serum to translocate from the nucleus to the leading edge of the cell, membrane ruffles and actin arcs where Flii colocalizes with Ras and GTPases 109.

The co-localization of Flii with Ras is consistent with findings using kinetic analysis based on competitive inhibition of Ras-dependent adenylyl cyclase activity which showed the association of the LRR domain of Flii with Ras 94,100. Ras proteins are involved in the regulation of the actin cytoskeleton where the interaction with Raf-1 connects actin to the mitogen-activated protein kinase (MAPK) pathway 100. Therefore, the interaction of Flii with Ras proteins provided a link to the MAPK pathway. The connection of Flii to a signaling pathway is further strengthened by a study which revealed the interaction of Flii to the phosphoinositol-signaling pathway in the regulatory events during ovulation 99.

In Swiss 3T3 fibroblasts, Flii located in the nuclear periplasmic region colocalizes with microtubules 109. In general, Flii protein localizes to actin based structures both in vitro and in vivo 94. The significance of this localization pattern remains unknown but suggests a possible function for Flii in actin regulation. However, the nuclear shuttling trait is not only limited to Flii as supervillin, CapG and gelsolin are also capable of nuclear translocation, and is the first indication that Flii may be involved in roles that are distinct from an actin remodeling protein.

35

1.4.2.5 Flii protein is a Negative Regulator of Wound Healing

Recent wound healing studies using mice heterozygous Flii knockout and mice over-

expressing the Flii gene demonstrated that Flii is an important negative regulator of the

wound healing process 110,111. The study showed that Flii regulates wound healing by

affecting cellular proliferation, motility and collagen I production in both epidermal

keratinocytes and dermal fibroblasts. Flii heterozygous mice displayed smaller and

more contracted wounds, coupled with an increase in α-smooth muscle actin-positive

myofibroblasts (Figure 1.7A). Conversely, Flii over-expressing transgenic mice

displayed increased wound area and dermal gape, reduced cell proliferation and

delayed epithelial migration (Figure 1.7C) compared to wild-type (Figure 1.7B) 111. It

was also found that Flii deficiency was associated with reduced collagen I production,

as evident from in vivo mouse wounds and in vitro siRNA knockdown fibroblasts. As

collagen I is the most highly produced collagen in response to wounding, the regulation

of it greatly affects the extent and outcome of cutaneous fibrosis and scar formation 112.

In burn wounds, pro-scarring TGFβ1 protein and gene expression were significantly

lower while anti-scarring TGFβ3 was higher in mice heterozygous for the Flii gene.

Furthermore, the addition of Flii neutralizing antibodies to in vitro incisional wounds

(Figure 1.7D) 111 as well as burn wounds 110 significantly improves wound healing, which indicated the possibility of modulating Flii activity to improve wound repair.

36

A

D

Flii+/-

B

WT

C

FliiTg/+

Figure 1.7

In vivo incisional wounding of WT, Flii heterozygous (Flii+/-) and Flii over-expressing mice (FliiTg/+) at day 3 post-wounding. (A-C) Incisional wounds showing improved healing in Flii+/- mice whereas FliiTg/+ mice show impaired healing compared to WT. (D)

Comparison of incisional wounds treated with Flii neutralizing antibodies in WT mice.

Images obtained from Cowin et al, 2007 111

37

1.5 TRANSFORMING GROWTH FACTOR BETA (TGFβ)

1.5.1 TGFβ superfamily

TGFβ is an important cytokine that is ubiquitously expressed in mammals and governs a substantial number of wound healing activities and processes in terms of cellular growth, differentiation, proliferation and adhesion. The TGFβ superfamily (Table 1.1) consists of highly conserved growth regulatory proteins that include bone morphogenenic proteins (BMP), activins, inhibins and Mullerian inhibitory factor (MIF) characterized by their 113. Here, we are focusing on three isoforms of TGFβ in mammals; TGFβ1, TGFβ2 and TGFβ3.

All TGFβ isoforms are expressed in the skin and play important roles in wound healing

114,115. In general, TGFβs act as chemoattractants during injury to direct an immune response towards the invading pathogens. TGFβ1 can stimulate the production of ECM such as collagen and therefore contribute to scarring. Studies using neutralizing antibodies to TGFβ1 and TGFβ2 showed improve healing with reduced scarring, indicating the detrimental effects of TGFβ1 and TGFβ2 isoforms in wound healing 116,117.

However, the exogeneous addition of TGFβ3 promoted wound healing indicating that

TGFβ3 is a positive regulator of wound healing 116,118.

38

TGFβs BMPs/GDFs Activins/Inhibins Others

TGFβ1 BMP-2 Activin A MIS TGFβ2 BMP-4 Activin B Lefty A TGFβ3 BMP-5 Activin C Lefty B BMP-6 Activin D BMP-7 Activin E BMP-8 BMP-9 BMP-10 BMP-15 Nodal GDF-1 GDF-3 GDF-5 GDF-6 GDF-7 GDF-9

Table 1.1

TGFβ superfamily members. The mammalian members of the TGFβ superfamily are sub-divided into (i) TGFβs, (ii) BMPs/GDFs (growth and differentiation factors), (iii)

Activins/Inhibins and (iv) other more distally related members, based on their structural characteristics.

39

TGFβs are potent cytokines with a range of regulatory receptors 119. The mature

peptides share 70-80% amino acid sequence identity and interact with the same

receptors 120,121. The regulation of TGFβ involves multiple checkpoints, from its

activation to signal transduction and these will be discussed below. Although extensive

research has been done on TGFβ, the full effects of TGFβ are still being elucidated as the

multifunctional nature of TGFβ is greatly influenced by a range of conditions including

cell type, growth conditions and presence of other growth factors.

1.5.1.1 TGFβ Activation and Structure

TGFβs are initially synthesized as large precursor polypeptides where they undergo

further proteolytic processing in the Golgi apparatus 119. The newly processed, mature

TGFβs non-covalently bind to the latency associated peptide (LAP) located at the N-

terminal domain, rendering it inactive by preventing interaction with its receptor. The

activation of TGFβ occurs via a multistep process involving different cell types,

proteolytic enzymes and the current micro-environment. There are many ways in which

TGFβ can be activated. These include acidification of microenvironments, action by

oxygen reactive species and steroid hormones such as anti-estrogens, retinoids, vitamin

D and glucocorticoids 122. However, during wound healing, TGFβ can be activated by secreted such as plasmin and thrombospondin which can cleave off the LAP.

40

In the wound site, reactive oxygen species and an acidic environment can both disrupt

the interactions between LAP and TGFβ to release bioactive TGFβ 123.

Another protein called latent TGFβ binding protein (LTBP) serves as a further control

during TGFβ activation. It covalently associates with LAP through two disulphide

bonds at the C-terminus 124. LTBP is important in the processing and secretion of mature

TGFβ such that its own expression is co-regulated with TGFβ 125. This also suggests a feedback loop between the expression of LTBP and TGFβ. Studies have shown that in the absence of LTBP, latent TGFβ secretion declines and it is mostly retained in the cis-

Golgi apparatus but, in the presence of LTBP, the association of LTBP with LAP

accelerates the secretion of mature TGFβ 126,127. This study has shown that LTBP

expression and TGFβ secretion are inter-connected and therefore serve as a regulatory

control over TGFβ activation.

The bioactive TGFβ released resembles a butterfly knot structure made up of two

12.5kDa molecules held together by four intrachain disulfide bonds formed by eight

cysteines. An additional sulphydryl bond on the last cysteine residue links another

monomer to form a 25kDa dimeric molecule (Figure 1.8 A, B). In most cases, TGFβs are

homodimeric molecules but heterodimeric TGFβ such as β1:β2, β1:β3 or β2:β3 can also

41

C

Adapted from Groppe et al, 2008 & Weiskirchen et al, 2009 128,129

42

Figure 1.8

Domain and three dimensional structure of mature TGFβs and interactions with TβRs.

(A) Within each TGFβ1 monomer, there are four intrachain disulphite bonds including

one conserved intrachain disulphite bond indicated as speckled line. TGFβ1 individual

monomers are linked together by a disulphite bond at position 77 of the mature

peptide. (B) The 4 intrachain disulphite bonds and the intermolecular disulphite bond

folds the dimer into a butterfly-like tertiary structure typical of all TGFβ isoforms. (C)

Ribbon and molecular surface representation of TGFβ ligand-receptor interaction in a ternary complex. TGFβ3 monomers are shown in blue and red. TβRI shown in yellow and TβRII shown in green.

43

be formed and this contributes to the extensive repertoire of biological processes

involving TGFβ 130. Mature TGFβ dimers will subsequently bind to their respective

target receptors to induce expression of downstream genes.

1.5.1.2 TGFβ Receptor Signaling

TGFβ signals transduce through heteromeric complexes of type I (65-70kDa) and type II

(85-110) serine/threonine kinase receptors (Figure 1.8C). These receptors are characterized by a cysteine-rich extracellular domain, a hydrophobic transmembrane domain and a C-terminal serine/threonine kinase domain. Functions of larger type III

(300kDa) receptors such as betaglycan and endoglin remain unclear but are not essential to the biological activities of TGFβ, instead they facilitates TGFβ interactions with the receptor signaling complex 131. It has been suggested that TGFβ receptors (TβR) are pre-formed even before binding to TGFβ ligand 132. The subsequent binding of the ligand causes a re-orientation rotation between the receptor chains, followed by simultaneous kinase activity. This may facilitate TGFβ signal transduction as a whole but its real purpose is not clear. However, this mechanism of receptor activation is also found in erythropoietin signaling pathways 133.

There are 11 different type I and six different type II receptors identified that bind to their respective ligands with different affinity. For example, TβRII has a higher affinity

44

of TβRII for TGFβ1 and TGFβ3 134,135. All type I (TβRI) and type II (TβRII) receptors have

similar structures and exist as homodimers in the absence of ligand. Upon binding to

TGFβ, TβRII recruits and phosphorylates serine and threonine residues in TβRI forming

an active heterotetrameric complex 136. This activated complex in turn initiates the

activation of downstream effectors, the Smad proteins (Figure 1.9).

1.5.1.3 Smad Effector Signaling

Smads are major TGFβ classical signaling transducers, mediating signals from the cell surface receptor to target genes in the nucleus (Figure 1.9). Smads are small molecular weight (42-60kDa) effectors. To date, there are a total of nine Smads that have been

characterized and divided into three groups based on their function. Receptor-activated

Smads (RSmads) that includes Smad 1-3, 5 and 8 make up the receptor-regulated Smad family and all consist of a conserved C-terminal SSXS motif. The other two groups are the common-partner Smad (co-Smad) that comprises Smad 4 and the inhibitory Smads,

6 and 7 137. All Smad proteins consist of two sequence homologies; Mad homology 1

(MH-1) at N-terminus and MH2 at C-terminus. Activation of the TGFβ receptor

heterotetrameric complex binds to RSmads 2/3 and phosphorylates amino acids SSXS

located at the C-terminus of RSmads which leads to a conformational change. Common

Smad 4 and inhibitory Smads 6 and 7 lack this SSXS motif. The phosphorylated RSmads

then dissociate from the receptor complex and form a heteromeric complex with co-

45

NOTE: This figure is included on page 46 of the print copy of the thesis held in the University of Adelaide Library.

Adapted from Derynck et al, 2003 138

46

Figure 1.9

Representation of TGFβ receptor and Smad signaling. At the cell membrane, TGFβ ligand binds to TβRII which then recruits and phosphorylates (P) serine and threonine residues in TβRI, forming an active heterotetrameric complex. The active heterotetrameric type I receptor consequently phosphorylate RSmads (Smad 2/3) which then associate with common Smad 4 and translocate into the nucleus where target gene transcription are regulated through additional interaction with DNA binding co-factors and other transcriptional factors. Phosphorylation of RSmads can be inhibited by Smads

6/7. X, DNA binding co-factor.

47

Smad (Smad 4) which then translocates into the nucleus and either directly or indirectly

associates with other transcriptional factors and binds DNA to induce target gene

expression 139. Rsmads 1, 5 and 8 also mediate signaling but more specifically further downstream of BMPs and their receptors.

Smad activities are also tightly regulated to serve as a layer of control for the signaling of TGFβ. It has been reported before that oncogenic Ras can inhibit the formation of

TGFβ induced Smad 2/3 complex as well as Smad 4 expression, therefore preventing nuclear translocation and accumulation to initiate gene expression 140. The stopping of

nuclear translocation can also be achieved by other proteins such as epidermal growth

factors (EGFs) which phosphorylates Smad linker regions to prevent nuclear

translocation. Smad can also be targeted by Smurf 1 (Smad uibiqutination regulatory

factor) which is an E3 ubiquitin ligase that recognizes and binds Smad 1 and Smad 5

(Zhu et al, 1999). Upon binding, Smurf 1 mediates the ubiquitination and subsequent

degradation of Smads which effectively arrests TGFβ signaling through the Smad

pathway.

1.5.1.4 TGFβs regulation in wound healing

TGFβ can also be regulated by TGFβ binding proteins such as the proteoglycan family

which includes decorin and fibromodulin that are also involved in wound healing by

48

regulating collagen formation and tensile strength 141,142. These small, leucine-rich

proteoglycans are components of the extracellular matrix which consist of a LRR motif

central protein core covalently linked to glycosaminoglycan side chains 143. These molecules belong to a distinctive protein class of the LRR superfamily 144. Decorin itself

interacts with TGFβ directly to modulate its activity 145 and a study 146 has substantiated

the fact reporting that decorin knockout cells have diminished TGFβ responsiveness.

Fibromodulin, another proteoglycan, also interacts with TGFβ to affect its activity. This

is of more relevance in wound healing because fibromodulin may potentially determine

the outcome of adult scarring by regulating TGFβ activity 142. Many members of the

proteoglycan family such as decorin, fibromodulin, biglycan and asporin directly

interact with TGFβ 147,148 via the LRR motifs 149. Therefore, it is likely that other proteins which contain LRR motifs are also able to bind to TGFβs.

The rate of TGFβ activation is also a major contributing factor to wound healing. As

previously discussed, proteases work to cleave off LAP to release mature TGFβ.

However, proteases do not work at the same efficiency. For example, matrix

metalloproteinase 2 and 9, are cell surface proteases which activates TGFβ and promote

tumor invasion and they work with different efficiency in activating TGFβ1, TGFβ2 and

TGFβ3 150. As a result, TGFβ1, TGFβ2 and TGFβ3 can be sequentially regulated at

different wound healing phases to affect the outcome of the wound. Furthermore, TGFβ

49

can be differentiated into two groups during blood clotting; the larger latent TGFβ complex (including LTBP, LAP and TGFβ) is released into the serum while the smaller latent TGFβ complex (LAP and TGFβ only) in the clot is activated gradually by proteases as the clot dissolves away 151. Therefore, this gradual release of TGFβ may be a means of maintaining TGFβ activity throughout the wound healing process.

1.6 Mitogen-activated Protein Kinase

1.6.1 The MAPK Signaling Pathways

The MAPK pathway is another common signaling pathway governing multiple cellular processes. The MAPKs are highly conserved protein enzymes responsible for the intracellular signaling cascades that transduce cell-surface receptor signals to the nucleus to induce target gene expression. The MAPK signaling pathways control many cellular processes such as cellular survival, proliferation, migration and adhesion which are also crucial during the wound repair process 152-154. Cell surface signals are transduced through the MAPK pathway through a three-tiered phosphorylation cascade, consisting of first tier MAPKK kinase, second tier MAPK kinase and the third tier MAPK 155. The main components of MAPK signaling cascades are shown in Figure

1.10. Given the complexity of the MAPK signaling cascade and to ensure specificity and maximize efficiency in order to prevent undesired activation of genes. Cells have to

50

NOTE: This figure is included on page 51 of the print copy of the thesis held in the University of Adelaide Library.

Adapted from Zhang et al, 2002 153

51

developed a mechanism known as protein scaffolding to regulate the MAPK pathway

156. The main function of protein scaffolds are to assemble multiple components of a particular MAPK signaling cascade such that it brings them to close proximity to operate efficiently. Protein scaffolds can also localize specific signaling molecules to a specific site in a cell and coordinate feedback signals to modify the MAPK cascade 157.

Additionally, protein scaffolds also protect activated signaling molecules from inactivation by other proteins 157. Using this method, cells are able to efficiently facilitate signal propagation from the cell membrane into the nucleus to induce target gene expression.

There are 5 distinctly regulated groups of MAPKs identified in mammals. These are the extracellular signal related kinase (MEK/ERK1/2), cJun N-terminal kinase (JNK), p38 proteins, ERK 3 and ERK 5 156, which are activated by their respective MAPKKs. For instance, MEK1/2 activates ERK1/2, MKK3/6 activates p38, JNKK1/2 activates JNKs and

MEK5 activates ERK5 158. More importantly, each second tier MAPKK can be activated by more than one first tier MAPKKK which greatly increases the complexity and dynamics of the MAPK signaling pathway 158.

52

1.6.1.1 ERK Pathway

The Raf/MEK/ERK cascade (reviewed in 156) is the most characterized pathway of the

MAPK signaling pathway which is mainly activated by protein tyrosine such as

the epidermal growth factor (EGF) receptor and the vascular endothelial

growth factor (VEGF) receptor 159. Ligand bound receptors undergo a conformational change that induces phosphorylation. Src homology 2 domain-containing proteins such as Brg-2 are then recruited to the membrane receptors and interact with the phosphotyrosine residues. Brg-2 becomes activated when recruited to the membrane and in turn activates Ras GTPase which will then hydrolyze guanosine triphosphate

(GTP) to guanosine diphosphate (GDP). GTP bound Ras consequently results in the activation of downstream effector proteins. The activation of Ras is regulated by two classes of proteins known as the GTPase activating proteins (GAPs) and guanine nucleotide exchange factors (GEFs) 160. GAPs reduce the availability of GTP-bound Ras by enhancing GTPase activity in Ras which increases the rate of GTP hydrolysis.

Conversely, GEFs facilitate the exchange of GDP for GTP therefore increasing the

numbers of GTP bound Ras. GTP bound Ras complex recruits and activates Raf at the

cell membrane which results in Raf catalyzing the phosphorylation of MEK1/2 which

subsequently activates ERK1/2 where they either continue activating downstream

targets or translocate into the nucleus to phosphorylate transcription factors 161.

53

1.6.1.2 p38Pathway

The mammalian p38 kinase consist of four members, p38α, p38β, p38γ and p38δ 162.

These p38 kinases can be activated by a range of extracellular stimuli such as UV

irradiation and cellular stress or by hormones and inflammatory cytokines. p38 kinases

are activated by MEK3 and MEK6 by phosphorylation. Once phosphorylated, p38 then

activates MAPK interacting kinases (Mnk1 and Mnk2).

1.6.1.3 JNK/SAPK Pathway

The cJun N-terminal kinase (JNK) family consist of three members, JNK1, JNK2 and

JNK3 161. JNK is activated by cytokines, growth factors or cellular stress. Upon

activation, JNK is phosphorylated by MEK4 or MEK7 and translocates into the cell

nucleus and further activates transcription factors including cJun, ATF2, STAT3 and

HSF1 161. Transactivated cJun results in increased expression of genes which promoters

consist of AP-1 sites such as TGFβ 153.

1.6.1.4 Phosphoinositide 3-kinase Signaling pathway

Phosphoinositide 3-kinase (PI3K) pathway (Figure 1.11) is another major signaling

pathway with many key regulatory roles in cellular processes including cellular

survival, proliferation and differentiation 163. PI3K/Akt is known to be involved in various TGFβ regulated processes including apoptosis and cell cycle arrest 164,165.

54

Ligand

Receptor Tyrosine Kinase

Ras PI3K LY294002

PIP2 PIP3

PTEN Akt

Response e.g cell proliferation

Figure 1.11

PI3K/Akt signaling pathway. Dotted lines represent activation of PI3K pathway via a

MAPK signaling protein Ras. Red lettering represent PI3K inhibitor.

55

PI3K are targets of receptor tyrosine kinases and G-protein coupled receptors as well as

Ras proteins. Integrins are also known to be able to activate the PI3K signaling pathway

166. Activated PI3K converts phosphatidylinositol-4-5-bisphosphate (PIP2) to phosphatidylinositol-3-4-5-triphosphate (PIP3) which subsequently activates Akt, a

serine threonine protein kinase 167. Regulation of the PI3K/Akt signaling pathway is

mediated by the tumour suppressor protein PTEN (protein phosphatase) which

dephosphorylate the 3' end of PIP3 and block Akt activation 167. It is also known that

Smad 3 is involved in TGFβ signal transduction and is a target of inhibition by

PI3K/Akt signaling cascade, which suggests the presence of cross-talks between the

signaling pathways.

1.7 Integration of MAPK and PI3k/Akt with TGFβ Signaling

Pathways

In any cellular signaling networks, cross-talk between related or different signaling

networks always exist as regulatory controls to achieve efficiency in cellular physiology.

As Ras signals through the MAPK pathway, the effects of TGFβ are mediated through

the Smad proteins. One convergence between these two signaling pathways is the

activation of p53 168. The Ras/MAPK cascade lies downstream of TGFβ receptor tyrosine

kinase (RTK) signaling. The activation of the RTK/Ras/MAPK signaling pathway

56

phosphorylates p53 which in turn interacts with TGFβ-activated Smads to target gene

expression. Studies have shown that ERK or JNK activation by RTKs can phosphorylate

Smad 3 at Thr 178, Ser 203 and Ser 207 protein residues in the linker region of Smad 3.

In addition, MAPKs can also regulate Smad 4 and Smad 7 by direct phosphorylation

140,169. In fact, ERK, JNK as well as p38 are all involved in the regulation of Smad 7170,171.

These findings indicated that TGFβ signaling pathway can be indirectly affected

through signaling cross-talk. On the other hand, PI3K/Akt signaling pathway can also

affect TGFβ signaling though its modulation of Smad 3. However, the regulation of

Smad 3 by the PI3K/Akt signaling pathway still remains unclear as it can either

potentiate or inhibit TGFβ responses172,173. It must be noted that TGFβ can also regulate

PI3K/Akt signaling, as evident by the increase in Akt activity in response to TGFβ 174.

1.8 RESEARCH AIM AND RATIONALE

Wound healing is undoubtedly complex with numerous molecular pathways affected

by a range of different factors. Multiple cross-talks exists within the molecular signaling pathways that are yet to be completely understood. This may contribute to the extensive control mechanism of the wound healing process hence providing a specific and effective approach in the regulation of wound healing. Whether or not all the different pathways converge together at some point remains to be elucidated.

57

Undisputedly, actin cytoskeletal proteins play a fundamental role during wound

healing, without which a lack of cell movement would lead to no wound healing. As a

result, it is very important that the actin cytoskeleton is regulated and this is

fundamentally achieved by the gelsolin family of actin remodeling proteins. In this

thesis, we describe two interconnected pathways affecting the cytoskeleton and wound

healing; first is the gelsolin androgen linked mediated pathway and second is the Flii

linked TGFβ signaling pathway. The underlining common denominator between these

pathways is the gelsolin family of actin remodeling proteins. This is highlighted by

recent findings on the role of gelsolin as a novel transcription regulator in AR-mediated wound healing. The implications of Flii in affecting TGFβ expression significantly expands the boundaries of potential roles for both gelsolin and Flii not just in the modulation of actin dynamics but also other biological roles.

In the first part of the thesis, we explored the relationship of gelsolin and the androgen receptor (AR) in androgen mediated wound healing. At present, steroid hormone mediated wound healing has been investigated and differences in wound repair has been observed. It is now known that hormone receptors can act as transcription factors which require co-regulators to activate target genes. This led to the identification of several co-regulators which includes gelsolin amongst other family members. Since gelsolin also plays important roles in wound healing, we hypothesize that gelsolin may

58

have a role in androgen mediated wound repair. The aim of this part of the thesis was

therefore to determine if gelsolin plays a part in androgen-mediated wound repair.

This thesis forms two main parts. The first section investigates the role of gelsolin in androgen mediated wound repair. Due to the inconclusive results obtained in part one of the thesis, the major focus of the thesis became the functional role of Flii in mediating

TGFβ gene expression and activity which forms the second part of the thesis.

TGFβ is a major cytokine during wound healing which functions to attract immune cells to prevent infection while also inducing gene expression. Flii, on the other hand, is a negative regulator of wound repair. Manipulation of Flii was observed to coincide with changes in TGFβ expression levels. As a result, our hypothesis was that Flii may be an important regulator of TGFβ expression. Therefore, in this part of the thesis, we aim to determine if Flii association with molecules provide a potential mechanism for modulating TGFβ gene transcription and signaling.

59

CHAPTER TWO

Role of gelsolin in androgen mediated wound healing

2.1 Introduction

2.1.1 Implications for the gelsolin superfamily in nuclear receptor

signaling

The expression of AR target genes inevitably requires the recruitment and interaction of

nuclear co-activators which cooperate together as a complex to initiate expression. The

identification of AR co-activators have revealed many different protein classes

including, β-catenin, breast cancer gene BRCA-1, cyclin E and more importantly, members of the gelsolin family 175. It was first reported in 2001 that supervillin from the

gelsolin family of ABP associates with the AR to enhance its transactivational activity

176,177. Since then, more members of the gelsolin family have been identified to be

involved in nuclear receptor signaling. Flii was also recently identified as a coactivator

in estrogen and thyroid hormone receptor transcription 105. More importantly, Gelsolin itself is identified to be a co-regulator for the AR 81. These findings implicate members of

the gelsolin family in the co-regulation of hormone nuclear receptor signaling. As a

result, there is a possible link for gelsolin in androgen mediated nuclear receptor

60

signaling during wound healing, given the relationship between gelsolin and AR in

which androgens also play an important part during the wound healing process.

2.1.2 Gelsolin and AR Nuclear Signaling

Looking more closely at the connection between AR and gelsolin, it has been previously

reported that gelsolin enhances the transcriptional activity of AR in the presence of an

agonist, either androgens or hydroxyflutamide 81. The evidence indicated that gelsolin

facilitates the nuclear translocation efficiency of AR during signaling which enhances

AR activity. Further investigation revealed two regions within AR in which gelsolin

interacts in the presence of a ligand; a central DNA binding domain and a ligand

binding domain in the COOH-terminal domain 81. Additionally, most AR co-regulators

have FXXMF or FXXFF peptide motifs which are present not just in gelsolin but also in

members of the gelsolin family including Flii, supervillin and advillin 178. This result

further implicates the involvement of the gelsolin family in nuclear receptor function.

Regardless, the enhanced transcriptional activity of AR is highly dependent on two factors, exposure to AR ligand and the co-expression of gelsolin. In fact, gelsolin has been demonstrated to colocalize with AR during nuclear translocation in the presence of ligand but had no effect if AR was absent 81. In other words, gelsolin remained in the

cytoplasm if AR expression is abrogated even in the presence of ligands. The gelsolin

61

behavior is expected and predictable quite simply because gelsolin lacks a nuclear

translocation signal. Therefore, it is only possible for gelsolin to co-translocate into the

nucleus while binding to other proteins which most likely in this case is AR. These findings suggest whilst the role of gelsolin in AR mediated gene expression may be to facilitate nuclear translocation, possibly through it's actin binding capabilities, its role in androgen mediated wound healing remains unknown.

Considering that Gelsolin plays an important role in cellular migration and androgens have an important role in the wound healing process, we sought to determine if gelsolin may be an important link between AR and androgen mediated wound healing.

62

2.2 Materials and Methods

2.2.1 Cells, Cell Culture

Human Foreskin Fibroblasts (HFFs) were cultured in Dulbecco’s modified Eagle’s

Medium (DMEM) supplemented with 10% fetal calf serum (FCS) and (100U

penicillin and 100ug/500ml streptomycin). Cell cultures were incubated at 37°C and 5%

CO2. Cells were serum starved in DMEM containing antibiotics for at least 3 hours or

otherwise stated prior to experimenting.

2.2.2 Scratch Assay

Cells were seeded onto sterile glass coverslips at a density of 2 x 105 cells/well, in a 6

well tissue culture plate and cultured until confluence. Cells were serum starved for 3

hours which were then linearly scratched multiple times using a P200 yellow pipette tip

producing approximately 20mm x 3mm wounds before treatment reagents were added.

Protein samples were obtained at time-points 0, 1, 3, 5, 10 and 24 hours after initial

wounding and analyzed. For immunocytochemical staining, samples grown on sterile

coverslips were fixed in ice-cold acetone for 10 seconds and placed in 1x PBS and stored

at 4°C.

63

2.2.3 Protein Extraction

For monolayer cell cultures, the cells were washed twice with ice-cold 1x PBS before

5ml of ice-cold 1x PBS was added and a cell scraper was used to gently lift the cells off.

The cell suspension was then centrifuged at 1000rpm for 5mins at 4°C and the supernatant discarded. 500µl of lysis buffer (50mM Tris pH 7.5, 1mM EDTA, 50mM

NaCl, 0.5% Triton-X100, protease inhibitor cocktail tablet (1 per 10ml - Complete Mini

(Roche, NSW, Australia) was then added to resuspend the cell pellet by pipetting up and down several times before incubating at 4°C for 30mins. The samples were then centrifuged at 14,000g at 4°C for 30mins and supernatant collected.

Bicinchoninic Acid (BCA) Quantification Kit (Pierce, Cat# 23227, Illinois, USA) was used as instructed by the manufacturer to quantify protein concentrations. Standards of

Bovine Serum Albumin (BSA) concentration from 20µg/ml to 2000µg/ml were prepared to obtain a standard curve against which the protein samples were quantified. Working reagent was made in the ratio of 50:1 solution A to solution B respectively. Both standards and protein samples were then transferred to 96-well microtitre plate at a volume of 10µl per well in duplicates which working reagent was then added. The samples were then incubated at 37°C for 30mins and absorbance measured at a wavelength of 570nm in a microplate reader.

64

2.2.4 Western Blotting

Protein samples were equalized by dilution with lysis buffer and loading buffer added before heating at 95°C. Protein fractions were then electrophoresed on a 10% separating

(3.35ml 30% Acrylamide-Bis Solution (37.5:1, 2.6% C, BioRad Laboratories, CA, USA),

1.25ml 3M Tris pH 8.9, 5.25ml distilled water, 125ul 10% SDS, 100ul Ammonium

Persulfate (APS) and 6.25ul TEMED (N,N,N’,N’ – Tetramethylethylene-diamine, Sigma

Aldrich, Sydney, Australia)) and 4% stacking (0.5ml 30% Acrylamide, 0.276ml 0.5M Tris pH 6.8, 4.104ml distilled water, 50µl 10% SDS, 40ul 10 % APS and 4µl TEMED) SDS-

PAGE gels at 100V for 90 mins and then transferred onto 0.2µm pore nitrocellulose membrane (Advantec MFS Inc, CA, USA) by wet transfer (Bio-Rad Laboratories,

Regents Park, NSW, Australia) using standard wet transfer - Towbin’s buffer (25mM

Tris, 192mM Glycine, 20% Methanol and 0.05% SDS) at 100V for 1 hour. Membranes were stained in Ponceau Red Staining Solution (Sigma Aldrich, Sydney, Australia) for 5 minutes to ensure equal protein loading and was later destained in distilled water and washed in PBS Tween (0.3% Tween/1x PBS).

Membranes were blocked in 5% milk blocking buffer (5% skimmed milk powder and

0.3% Tween-20 diluted in 1x PBS) for 1 hour and primary antibodies diluted in blocking buffer added to the membrane and incubated overnight at 4°C. Stringent washes with blocking buffer were performed every 15 minutes for an hour before appropriate

65

secondary antibodies conjugated with horse radish peroxidise (HRP) was added for 1

hour at room temperature. Washes were then performed before signal detection using

Super Signal West Femto Maximum Sensitivity Substrate (Pierce Biotechnology,

Rockford, USA) and signal capture using GeneSnap analysis program (Syngene,

Maryland, USA). Membranes were stripped and re-probed with β-tubulin as a loading control (Sigma Aldrich, Sydney, Australia).

2.2.5 Immunocytochemistry

Sterile glass coverslips were placed in six well culture plates and cells were seeded at a

5 density of 3 x 10 cells per well and cultured in an incubator at 37°C with 5% CO2. Cells

were serum starved for 3 hours before adding treatments and incubating for a further

30 minutes. This was followed by fixing cells with cold acetone for 10 seconds and

placed back into 1x PBS containing six well plates. Washes were performed after every

treatments using 1x PBS. 3% Normal Horse Serum (NHS) diluted in 1x PBS were used

to block cells for 30 minutes at room temperature before incubating with mouse anti-AR

antibodies overnight at 4°C(Refer to Table 2.1). Secondary anti-rabbit Alexa Fluor 488

were added for 1 hr in the dark at room temperature (Refer to Table 2.1). This was then

followed by incubating with mouse anti-gelsolin antibodies for 1 hr at room

temperature. The respective biotinylated secondary antibodies were added and

coverslips incubated for 1 hr in the dark at room temperature. After this, streptavidin

66

Concentration used Antibody Manufacturer Catalog # Raised in Western Immunostaining

Androgen Receptor Santa Cruz sc815 Rabbit 1µg/ml 2µg/ml

Gelsolin BD Biosciences 610412 Mouse 1µg/ml 2µg/ml

β-Tubulin Sigma T4026 Mouse 0.01µg/ml N/A

HRP-conjugated DAKO P0448 Goat N/A 2µg/ml 2° Anti-Rabbit IgG HRP-conjugated DAKO P0447 Goat N/A 2µg/ml 2° Anti-mouse IgG 2° Anti-rabbit Invitrogen A11008 Goat N/A 2µg/ml Alexa Fluor 488 Biotinylated Vector BA2000 Horse N/A 2µg/ml 2° Anti-mouse IgG Laboratories Strepdavidin Invitrogen S32355 N/A N/A 2µg/ml Alexa Fluor 555

Table 2.1

Information of antibodies used in western analyses and immunofluorescent staining.

67

conjugated Alexa Fluor 555 was added for 1 hr in the dark at room temperature. DAPI

was added to the cells before a final wash and then mounted onto a slide

using DAKO Fluorescent Mounting Medium (DAKO, Botany, Australia). Slides were stored in the dark at - 20°C. Integrated fluorescence intensity was determined using

AnalySIS software (Soft-Imaging System GmbH, Munster, Germany). Negative controls

were included to demonstrate antibody staining specificity. Control samples undergo

the exact staining procedure omitting either the primary or the secondary antibody. All

control samples had negligible immunofluorescence.

2.2.6 siRNA knockdown Assay

HFFs were seeded into 6 well tissue culture plates and cultured overnight to achieve

30% to 50% confluence at time of transfection. Small interfering RNA (siRNA) for

gelsolin (M-007775-01, Dharmacon, CO, USA) and negative control siRNA (Cat#4611,

Ambion, TX, USA) were used to knockdown gelsolin expression levels in HFFs. siRNA

were transfected into the cells using Lipofectamine 2000 (Invitrogen, Carlsbad, USA).

Both siRNA and Lipofectamine 2000 were diluted in Opti-MEM I Reduced Serum

Medium (Invitrogen, Carlsbad, USA). 250µl of siRNA (optimized to 100nM per well)

was mixed with 4µg of Lipofectamine 2000 diluted in 250µl Opti-MEM and were

allowed to complex at room temperature for 20 minutes. 500µl of siRNA:Lipofectamine

2000 complex was then added to each well, mixed and cells incubated for 6 hours before

68

replacing transfection media with DMEM containing 10% FCS only. Cells were incubated for 24-48 hours prior to gene knockdown assessment by real-time quantitative PCR and western blot analysis.

2.2.7 RNA Extraction

RNA extraction from cell cultures only required scraping cells from culture flasks after adding Trizol Reagent (Invitrogen, Victoria, Australia). Samples were transferred into fresh eppendorf tubes and centrifuged at 12,000g at 4°C for 10mins to remove cell debris. The samples were incubated for 5mins at room temperature before adding 200µl of chloroform to each tube and were mixed thoroughly by hand for 15secs. The samples were kept at room temperature for 3mins and centrifuged at 12,000g for 15mins at 4°C.

The aqueous phase containing RNA was transferred into a fresh tube and 500µl of isopropanol added to precipitate the RNA. The samples were then incubated at room temperature for 10mins before centrifuging at 12,000g for 10mins at 4°C. The supernatant was discarded and the residual pellet washed with 1ml of 75% ethanol.

Finally, samples were centrifuged at 7500g for 5mins at 4°C and supernatant discarded.

The pellet was dried and re-dissolved in 50µl DEPC water.

69

2.2.8 DNase Treatment and RNA Quantitation

RNA samples were subjected to DNA-free DNase Treatment and Removal Kit

(Ambion, TX, USA) as instructed by the manufacturer to remove any contaminating

genomic DNA. Firstly, RNA samples were treated with 0.1 volume of 10x DNase Buffer and 1µl of DNase I and incubated at 37°C for 30mins. Following this, 0.1 volume of

DNAse Inactivating Reagent was added to the samples and incubated at 2mins at room temperature with occasional mixing. The samples were then centrifuged at 10,000g for

90secs and supernatant transferred to fresh tubes. RNA was quantitated by diluting 1 in

20 with RNase free water and 100µl duplicates were quantified using a Pharmacia

Biotech GeneQuant RNA/DNA Calculator using RNase-free water as a blank.

Absorbance at 260nm and 280nm were measured that quantify RNA absorbance as

µg/µl concentration. Purity of RNA was confirmed by the A260/A280 ratio and a value

between 1.7 to 2.0 indicates good RNA quality.

2.2.9 Complementary Deoxyribonucleic Acid (cDNA) Synthesis

cDNA was synthesized from RNA using reverse transcription. Each reaction contains

1µg of RNA with 4µl 2.5µM dNTPs (dATP, dCTP, dGTP and dTTP, 100mM each,

Promega, WI, USA) and 2µl Oligo(dt)12-18 Primer (25µg at 0.5µg/µl, Invitrogen, Victoria,

Australia). This was heated at 85°C for 3mins and placed in ice immediately. 2µl 10x

70

Stratascript Buffer (Stratagene, Epson, UK), 1µl RNasin (Promega, WI, USA) and 1µl

Stratascript Reverse Transcriptase (Stratagene, Epson, UK) were added to the mixture

and heated at 42°C for 60mins followed by 92°C for 10mins before cooling it down on

ice. A control sample was prepared with the reagents above with the exclusion of

reverse transcriptase for use in the Real-Time quantitative-Polymerase Chain Reaction

(RTq-PCR) as a negative control.

2.2.10 Real-Time quantitative-Polymerase Chain Reaction (RTq-PCR)

Each PCR reaction tube containing cDNA was set up to a final concentration of 1x SYBR

Green, 1x Amplitag PCR buffer, 3mM MgCl2, 5mM dNTPs, 0.9µM primers (forward

and reverse), 1.25 Units of AmpliTag Gold DNA polymerase in 25µl of H2O. The primer

sequences were generated in Genbank and were as follows; Gelsolin (forward) 5’ CAG

ACA GCC CCT GCC AGC ACC C 3’ and (reverse) 5’-GAG TTC AGT GCA CCA GCC

TTA GGC-3’. Cyclophillin A (forward) 5’-GGT TGG ATG GCA AGC ATG TG-3’ and

(reverse) 5’-TGC TGG TCT TGC CAT TCC TG-3’.

2.2.11 Proliferation Assay

WST-1 proliferation reagent (Cayman Chemical, WI, USA) was used to assess the rate of

proliferation of HFFs and primary cell cultures. This is a tetrazolium salt based reagent

71

that converts to soluble formazan by dividing cells therefore directly correlates to

cellular metabolic activity and their proliferation rate. Cells were seeded into 96-well

microtitre plates at a density of 105 cells per well in 100µl of 10%FCS DMEM with antibiotics. Cells were then incubated at 37°C CO2 incubator overnight and serum

starved for 3 hours the following day. Cells were then treated with a range of

treatments and incubated for 48 hours or otherwise stated and assayed by adding 10µl

of WST-1 reagent and mixed thoroughly for 1 min on an orbital shaker. The cells were

then incubated at 37°C for 1 hour as per manufacturer’s protocol before measuring dual

absorbance at 450nm and 600nm on a Tecan microplate reader.

2.2.12 Wound Closure Assay

Cells were seeded into 6 well culture plates and cultured until confluent. Cells were

serum starved for 3 hours prior to wounding and addition of treatments. Using a P200

pipette tip, a circular wound was created. Photographs were taken at time-points 0, 24,

36, 48, 60, 72 and 84 hours after initial wounding. Rate of wound closure was

determined, which is defined by the residual area and expressed as a percentage of the

initial wound size.

72

2.2.13 Statistical Analysis

All statistical differences were determined using the Student’s t-test or ANOVA. P value of less than 0.05 was considered significant.

73

2.3 Results

2.3.1 Gelsolin and AR responses to wounding in vitro

To determine if gelsolin and AR were affected by wounding, HFFs were subjected to an

in vitro scratch wound assay. Figure 2.1 shows protein extracted at 0, 1, 3, 5, 10 and 24

hours post-wounding and analyzed using western analysis. Both gelsolin and AR

protein expression were increased in response to wounding and had a similar temporal

expression pattern during the 24 hour time-course, albeit a difference in expression peak times (Figure 2.1). Gelsolin expression increased and peaked at 3 hours post- wounding before declining back to basal levels after 24 hours. Similarly, AR expression also increased in response to wounding but AR expression had a lag period before increasing at 3 hours and peaked at 5 hours post-wounding before declining back to baseline levels.

The expression pattern and localization of gelsolin and AR were compared. Cellular localization of gelsolin and AR was determined using immunofluorescence staining

(Figure 2.2). Here, it is shown that AR expression increases in response to wounding.

This is reflected by the increased staining intensity at and after 3 hours post-wounding

(Figure 2.2D, G, J, M) compared to time 0 (Figure 2.2A). During the 24 hour wounding

74

A IB: 0 1 3 5 10 24 B IB: 0 1 3 5 10 24 Gelsolin AR 88 kDa 110 kDa

β-tubulin β-tubulin 55 55 kDa 55 kDa

C D

300 230 * 210 * 250 190 * * 170 * 200 * * * 150 130 150 110 Percentage Change Percentage Change Percentage 100 90 70 50 50 0 1 3 5 10 24 0 1 3 5 10 24 Hours Hours

75

Figure 2.1

Western analyses of gelsolin and AR expression in time-course wounding in HFFs.

(A,C) SDS-PAGE. Confluent HFF monolayers were scratched multiple times and total lysate collected at 0, 1, 3, 5, 10 and 24 hours post-wounding and immunoblotted with anti-gelsolin and anti-AR antibodies, giving full length 88 kDa gelsolin and 110 kDa AR protein band respectively. β-tubulin was used as a loading control. Both gelsolin and

AR expression were increased in response to wounding. (B,D) Quantitation of gelsolin and AR protein bands normalized to β-tubulin band and expressed as a percentage to unwounded control (0 hour). Gelsolin expression peaked at 3 hour post-wounding compared to AR which peaked at 5 hour post-wounding before declining back to basal after 24 hours. n = 4, *p < 0.05 compared to time at 0 hours.

76

time-course, AR expression peaked at 5 hours after wounding before gradually

decreasing at 10 and 24 hours post-wounding. AR staining was also detected in the

cytoplasm and nucleus of both unwounded and wounded HFFs (Figure 2.2A, D, G, J,

M). Gelsolin expression also increased in response to wounding. Upon wounding at

time 0 hour, gelsolin expression was low (Figure 2.2B). The highest gelsolin expression

was observed at 3 hours post-wounding (Figure 2.2E) before decreasing at 5, 10 and 24

hours (Figure 2.2 H, K, N) after wounding. Gelsolin expression was detected in the

cytoplasm and nucleus of both unwounded and wounded HFFs.

Co-localization between AR and gelsolin was also investigated by merging identical

field of view images of AR and gelsolin staining to create composite images (Figure

2.2C, F, I, L, O). Cytoplasmic and nuclear co-localizations are represented by arrowheads and arrows in the merged images respectively. Co-localization of AR and gelsolin were observed throughout the wound repair time-course. Strong co- localization of AR and gelsolin were observed at and after 3 hours post-wounding

(Figure 2.2F, I, L, O). Although cytoplasmic co-localization was observed, the majority

of co-localization was in the nucleus and the nuclear periplasmic region (arrows).

77

AR Gelsolin Merge w B w C w A

0 Hour

D w E w F w

3 Hours

G w H w I w

5 Hours

J w K w L w

10 Hours

M N O

24 Hours w w w

78

Figure 2.2

HFFs cultured on sterile glass coverslips were wounded and fixed in cold acetone at 0,

3, 5, 10 and 24 hours post- wounding. Wound (w) is defined by the dotted lines Red

represents AR staining (A, D, G, J, M), green represents gelsolin staining (B, E, H, K, N)

and blue represents nuclear staining. Merge images are composite images of AR and

gelsolin staining (C, F, I, L, O). Cytoplasmic co-localizations (arrow heads) and nuclear co-localizations (arrows) are represented by areas of yellow color staining. n = 6, w = wound. Scale bar = 100µm in (O).

79

2.3.2 Effects of DHT addition in gelsolin and AR localization

To study the effects of adding the active metabolite DHT on AR and gelsolin expression

in HFFs, immunofluorescent staining was used. DHT is the more active metabolite of

the hormone testosterone and is a ligand to the AR. A physiological concentration of 10-

8M DHT was added to HFFs for a period of 24 hours before immunostaining with anti-

gelsolin and anti-AR antibodies. In the absence of DHT, AR and gelsolin expression was

relatively subdued and localized predominantly to the cytoplasm (Figure 2.3A, B). Little

AR and gelsolin staining was observed in the cell nucleus. Co-localization of AR and

gelsolin was investigated in untreated cells. In the absence of DHT, AR and gelsolin

were predominantly co-localized to the nuclear peripheral region (arrowhead) in the

cytoplasm (Figure 2.3C). Addition of DHT on the other hand resulted in significantly

higher staining intensities for AR and gelsolin (Figure 2.3D, E). The increased

expression of AR and gelsolin were predominantly localized to the nuclear periplasmic

region in the cytoplasm, in addition to the cell nucleus, indicating partial translocation

of both proteins into the cell nucleus. Interestingly, AR and gelsolin co-localization were

observed in the cell nucleus (arrows) in addition to the nuclear peripheral region seen

for the untreated cells (Figure 2.3F).

To investigate the effects of DHT addition on AR and gelsolin expressions, three doses

80

AR Gelsolin Merge A B C

Untreated

D E F

+DHT

Gelsolin Fluorescent Intensity in HFFs Treated with AR Fluorescent Intensity in HFFs Treated with DHT G DHT H 200 * 140 * 180 * * * 120 160 140 100 120 80 100 60 80 60 40

Fluorescent Intensity 40 Fluorescent Intensity 20 (% Normalized Control)to 20 (% Normalized to Control) 0 0 Untreated 10-9M 10-8M 10-7M Untreated 10-9M 10-8M 10-7M

81

Figure 2.3

Effects of DHT addition on gelsolin and AR expression and localization in HFFs.

Confluent HFFs were serum-starved then treated with 10-8M of DHT in (A-F) and

varying concentrations of DHT as shown respectively in (G,H). Dual

immunofluorescent staining was then performed on HFFs treated with DHT. Red

represents AR staining (A, D), green represents gelsolin staining (B, E), and blue

represents nuclear staining. Merge images are composite images of AR and gelsolin

staining (C, F). Cytoplasmic co-localizations (arrow heads) and nuclear co-localizations

(arrows) are represented by areas of yellow color staining. n = 3 (A-F), n = 6 (G, H).

*p<0.05 compared to untreated control. Scale bar = 20µm in (F).

82

of DHT, low (10-9M), physiological (10-8M) and high (10-7M) concentrations were added

to HFFs for a period of 24 hours. The cells were then stained with either AR or gelsolin

antibodies and their fluorescent intensities quantitated. DHT addition slightly increased

gelsolin expression at 10-8M and 10-7M concentrations (Figure 2.3G). DHT at 10-9M

concentration did not significantly affect gelsolin expression as compared to the

untreated cells. On the contrary, DHT significantly increase AR expression in HFFs. All

concentrations of DHT at 10-9M, 10-8M and 10-7M increased AR expression by a

minimum of 50% (Figure 2.3H). No differences were observed between the

concentrations suggesting that AR is highly sensitive to the presence of DHT even in

small amounts.

2.3.3 Time-course localization study of gelsolin and AR in the presence

of DHT.

A time-course localization study of gelsolin and AR was conducted using

immunofluorescent staining. Since nuclear receptor signaling is a rapid response

pathway, it would be expected to show differences in localization of both gelsolin and

AR. The effects of DHT treatment on short-term and long-term localization of gelsolin and AR were investigated. Short-term time-points include 0, 5, 10, 15, 30 and 60 minutes

(Figure 2.4). Long-term time-points include 2, 5, 10, 24, 48 and 72 hours (Figure 2.5).

83

AR Gelsolin Merge A B C

0 Min

D E F

5 Mins

G H I

10 Mins

J K L

15 Mins

M N O

30 Mins

P Q R

60 Mins

84

Figure 2.4

HFFs cultured on sterile glass coverslips were fixed in cold acetone at 0, 5, 10, 15, 30 and

60 minutes after the addition of 10-8M DHT. Red represents AR staining (A, D, G, J, M,

P), green represents gelsolin staining (B, E, H, K, N, Q) and blue represents nuclear staining. Merge images are composite images of AR and gelsolin stainings (C, F, I, L, O,

R). Cytoplasmic co-localizations (arrow heads) and nuclear co-localizations (arrows) are represented by areas of yellow color staining. n = 6. Scale bar = 100µm in (R).

85

At 0 minutes, co-localization between gelsolin and AR was visible but largely localized

in the cytoplasm (Figure 2.4A-C). Nuclear localization of both gelsolin and AR was also

observed. The reason that nuclear staining was seen was due to the time lag which

occurs when DHT was added to the time the image was taken. As a result, the effects of

DHT had already begun by the time the image was taken, therefore explaining the

observation of nuclear staining at time 0. However as time progressed, both gelsolin

and AR translocated into the nucleus gradually, where staining of both proteins was

either concentrated in or around the nucleus. At 5 minutes, co-localization was

observed clearly in the nucleus, represented as pink which becomes increasingly more

defined as time progressed (Figure 2.4D-F). In addition, at 10 minutes post-treatment, staining for AR were beginning to increase significantly, indicating increased expression (Figure 2.4G-I). At 15 to 60 minutes, localization of gelsolin and AR were not significantly different than at 10 minutes except that at 60 minutes, staining for AR was predominantly nuclear where it also co-localized strongly with gelsolin (Figure 2.4J-R).

At longer time-points, gelsolin and AR expression appeared to be increased compared

to time 0 minutes (Figure 2.5A-R). Strong nuclear staining was observed for AR as well

as gelsolin. Gelsolin staining was also distributed throughout the cytoplasm, in addition

to the nucleus. At 5 hours post-treatment, HFFs appeared to be denser compared

86

AR Gelsolin Merge A B C

2 Hrs

D E F

5 Hrs

G H I

10 Hrs

J K L

24 Hrs

M N O

48 Hrs

P Q R

72 Hrs

87

Figure 2.5

Gelsolin and AR long-term response to the addition of 10-8M DHT. (A-R) HFFs were

treated with 10-8M DHT and fixed in acetone at 2, 5, 10, 24, 48 and 72 hours after

treatment and then stained with respective antibodies to gelsolin and AR. Red

represents AR staining and green represents gelsolin staining. Longer treatment with

DHT did not yield significant differences in gelsolin and AR localization. At 2 hours

post-treatment and beyond, gelsolin staining and AR staining were localized predominantly to the nucleus where they co-localized as shown in the merge column.

No translocation from the nucleus to cytoplasm was observed after an extended period.

88

to the earlier timepoints (Figure 2.5D-F). No other significant effects were observed after

5 hours of treatment with DHT. Regardless, in the presence of DHT, AR staining remains to be strongly localized to the nucleus. Gelsolin on the other hand was most strongly localized to the nucleus at 60 minutes post-treatment but was relatively evenly distributed throughout the cell at other time-points.

Using multiple images gathered from immunofluorescence microscopy, the staining intensities of gelsolin and AR at each time-point were quantitated and graphed in

Figure 2.6. As expected, during the early time-points from 0 to 60 minutes, no increase in staining was observed, which indicated that expression of both gelsolin and AR remained relatively constant when treated with DHT in the first 60 minutes (Figure 2.6

A,B). Looking at longer time-points, an increase in staining intensity was observed for gelsolin at and beyond 2 hours, indicating that expression levels were increasing from 2 hours of treatment with DHT (Figure 2.6C). Expression continued to increase the past fifth hour, peaking at 48 hours, where expression levels were approximately 35% higher than at time 0, before leveling off at 72 hours. AR on the other hand also had a similar pattern of increased expression (Figure 2.6D). Expression was significantly higher at 10 hours and continued to increase before leveling off at 72 hours. AR expression was approximately 60% higher than at time 0.

89

120 120

100 100

80 80 60 60 40 40

20 20 (%) Intensity AR Staining Gelsolin Staining Intensity (%) 0 0 0 5 10 15 30 60 0 5 10 15 30 60 A B Time (mins) Time (mins)

160 200 * * * 140 * * 180 * 160 * 120 * 140 100 120 80 100 60 80 60 40 40

20 (%) Intensity AR Staining 20

Gelsolin Staining Intensity (%) 0 0 0 2 5 10 24 48 72 0 2 5 10 24 48 72 C D Time (Hours) Time (Hours)

Figure 2.6

Multiple images obtained from immunofluorescence microscopy in figure 2.4 and

figure 2.5 were analyzed. AR and gelsolin staining intensities were quantitated and

results normalized to time 0 hours and expressed as a percentage. (A,B) Short time- course treatment with DHT (0, 5, 10, 15, 30, 60 minutes). (C,D) Long time-course treatment with DHT (0, 2, 5, 10, 24, 48, 72 hours). n = 6 per time-point. *p<0.05 compared to Time 0 hours.

90

2.3.4 Gelsolin gene knockdown using silent RNA technology

To determine whether gelsolin plays a role in the androgen mediated wound repair

process, functional analyses by reducing gelsolin expression were done. Silent RNAs or

short interfering RNAs (siRNA) were used to knockdown gelsolin expression levels.

Optimization of the technique was carried out prior to knocking down gelsolin

expression (results not shown). Three different concentrations of gelsolin siRNA, 50nM,

100nM and 150nM were used to show that 50nM was sufficient to significantly

knockdown gelsolin expression by more than 85% (Figure 2.7A). Additional siRNA did

not further decrease gelsolin expression. Therefore 50nM of siRNA were used in all

subsequent gelsolin knockdown experiments. After knockdown, gelsolin expression

remained significantly lower compared to the control (Figure 2.7B). Expression

remained at approximately 18% and 21% at 72 hours and 92 hours post-transfection respectively. To further illustrate gelsolin siRNA specificity, a non-specific scrambled siRNA sequence was used to transfect HFFs and the end result showed no significant differences in gelsolin expression compared to the control (Figure 2.7C). However, it was observed that decreasing gelsolin expression had an effect on AR expression as shown in figure 2.7D. Specifically, AR expression increased when gelsolin expression was reduced.

91

120 120 100 100

80 80

60 60

40 40 * * 20 * 20 Gelsolin Expression (%) * * Gelsolin Expression (%) 0 0 A Control 50nM 100nM 150nM B Control 72 Hours 92 Hours

1.2 350 * 300 1 250 0.8 200 0.6 150

Fold Change 0.4 100 AR Expression (%) 0.2 50 0 0

C Control Scrambled siRNA D Control 50nM 100nM

Figure 2.7

Gelsolin expression was knocked down using silencing RNA technology. HFFs were

treated with gelsolin specific siRNA to decrease gelsolin expression. mRNAs were

collected and levels assessed after 48 hours using RT-qPCR. (A) Gelsolin expression levels using 50nM, 100nM and 150nM of siRNA. (B) Gelsolin expression remained suppressed after 92 hours from initial siRNA treatment. Expression level at 92 hours was approximately 20% of control. (C) Using a non-specific, scrambled siRNA sequence, gelsolin expression showed no significant changes. (D) The effects of gelsolin knockdown caused AR expression to increase significantly. n = 3. *p<0.05 compared to control.

92

2.3.5 Response to DHT treatment in gelsolin-ablated HFFs

The effects of DHT addition on HFFs proliferation were determined by adding

increasing concentrations of DHT (10-7M, 10-8M and 10-9M) and assessing the effect

using the WST-1 colorimetric assay (Figure 2.8). As expected, DHT increased HFF

proliferation by more than 50% at 10-8M DHT compared to the control (Figure 2.8A). 10-

7M and 10-9M DHT also increased proliferation but by a smaller amount. In gelsolin-

ablated HFFs, proliferation was significantly lower compared to control (Figure 2.8B).

This highlighted the importance of gelsolin in cellular proliferation. However, when

DHT was added to gelsolin knockdown HFFs, proliferation increased to be comparable

to the control. 10-7M DHT treated HFFs remained significantly lower than that of the

control (Figure 2.8M). Although proliferation had risen, DHT did not increase

proliferation as significantly as that of HFFs with normal gelsolin levels.

The DHT effects on HFF migration were also investigated using a wound assay that

measures rate of closure over 0, 24, 36, 48, 60, 72 and 84 hours (Figure 2.9). Addition of

DHT at 10-7M, 10-8M and 10-9M all enhanced wound closure (Figure 2.9A). Gelsolin is an

important actin regulator and plays a significant role in cellular migration. As a result, when gelsolin levels were decreased following siRNA transfection, HFF rate of wound closure fell considerably. Addition of DHT at 10-7M and 10-9M concentrations did not

93

160 * * 140 * 120 100 80 60 40 Proliferation (% to control) (% to Proliferation 20 0

A Control 10-9M DHT 10-8M DHT 10-7M DHT

120

100 * 80 *

60

40

20 Proliferation (% to control) (% to Proliferation

0 Control Gelsolin KD Gelsolin KD + Gelsolin KD + Gelsolin KD + B 10-9M 10-8M 10-7M

Figure 2.8

Proliferation of HFFs treated with gelsolin siRNA and DHT. HFFs were serum starved for a minimum of 3 hours before treatments. The assay is based on WST-1 colormetric detection of metabolic enzymes and measured using a spectrophotometer. (A)

Treatment of DHT (10-9M, 10-8M, 10-7M) on normal confluent HFFs (B) Treatment of

DHT (10-9M, 10-8M, 10-7M) on gelsolin siRNA knocked down HFFs. DHT was added to

HFFs 48 hours after initial gelsolin siRNA transfection. n = 6, *p<0.05 compared to controls.

94

100 Control *a-b 10-9M 80 10-8M *c-e 60 10-7M

40 *f-h Wound SizeWound (%) *i-k 20

0 0 24 36 48 60 72 A Time (Hours)

100 *a-h Control *i-o Gelsolin KD 80 Gelsolin KD + 10-9M

Gelsolin KD + 10-8M 60 Gelsolin KD + 10-7M *p-u

40

Wound SizeWound (%) *v-z 20

0 0 24 36 48 60 72 84

B Time (Hours)

95

Figure 2.9

In vitro wound closure assay. HFFs were serum-starved before wounding and measurements recorded at 0, 24, 36, 48, 60, 72 and 84 hours. Wound closure was measured by taking wound size area and converting to percentage of initial wound size. (A) Effects of DHT addition on rate of wound closure. The treatment of DHT accelerated the rate of wound closure as compared to the controls. Rate of wound closure had no differences between DHT concentrations. (B) Effects of gelsolin ablation on HFF wound closure. Wound closure in gelsolin knockdown HFFs was hindered considerably indicating the importance of gelsolin in cell migration. The 84 hours time- point was included to show complete wound closure. Treatment with 10-9M DHT had

no significant effects compared to gelsolin knockdown HFFs and rate of wound closure

was significantly impaired compared to the control. The addition of 10-8M DHT

however did increase the rate of wound closure but remained slower compared to the

control. Addition of 10-7M did increase rate of wound closure slightly at the first 36 hours of wound closure but was of no difference at 48 to 84 hours post-wounding. n = 4,

p<0.05

96

change the rate of closure, it remained similar to gelsolin siRNA treatment group.

Although in the early time-points, 10-7M DHT treated HFFs had better closure, the significance was low and was identical to gelsolin knockdown cells towards the later time-points. Conversely, gelsolin knockdown HFFs treated with 10-8M DHT had an

improved rate of wound closure. However, it was not enough to recover to that of the

control group. Even with addition of DHT, the rate of wound closure in gelsolin

knockdown HFFs was still notably decreased.

97

2.4 Discussion

Defining the relationship between two different proteins, gelsolin and the AR is integral

to further understand the fundamentals of hormone mediated wound healing and

potentially could lead to the development of rational pharmacogenomic based

therapeutics. This study has attempted to characterize the relationship between gelsolin

and AR during in vitro wound healing.

Initial studies were focused on elucidating both the gelsolin and AR expression pattern during an in vitro scratch wound assay. The results from this experiment showed up-

regulation of both gelsolin and AR in response to wounding. More specifically, it

showed a temporal increase of gelsolin and AR expression, peaking at 3 and 5 hours

respectively. This increase in gelsolin expression is consistent with findings from

studies showing motility defects, including pronounced stress fibers in skin fibroblasts

from gelsolin knockdown mice 78. The AR increase in expression was also consistent

with the full time-course wound repair findings by Ashcroft et al, 2002. Furthermore,

quantitation of immunofluorescent staining intensities also supported the findings,

showing up-regulation of gelsolin and AR during wound healing. By using a pure

population of human foreskin fibroblasts (HFF), the localization of both gelsolin and AR

could be visualized. Initially, expression of gelsolin was relatively low and localized

98

mainly to the cytoplasm, as with AR. As time progressed, staining for both gelsolin and

AR increased before leveling off. Nuclear staining was also observed. Since gelsolin does not have a nuclear localization signal, its presence in the nucleus will most likely be serving as a co-regulator for AR target genes 81. During wound healing, gelsolin expression was highest at the leading edge of migrating cells. This is expected as actin cytoskeletal structures such as filopodia have to be regulated for cellular migration 179.

AR was also up-regulated in the leading edge cells which suggests a role for AR in the wound healing process. Since DHT was absent in the cell populations, the increase in

AR would have to be induced by an alternate pathway and maintained by a positive feedback loop 175.

Stimulation of HFFs using a dose response of DHT showed partial nuclear translocation of both gelsolin and AR. Results from Nishimura et al, 2003 showed complete nuclear translocation using COS-1 cells treated with either DHT or hydroxyflutamide. This is in contradiction to this study where no complete nuclear translocation of gelsolin and AR has been observed throughout the time-course experiment using HFFs. Strongest nuclear staining occurred within the first hour of DHT stimulation. This discrepancy may be attributed to the difference in cell line and most likely the sensitivity of cells to

DHT stimulation. Regardless, the co-translocation of both gelsolin and AR in HFFs would suggest cooperative association to initiate AR response genes for cellular growth.

99

AR expression was induced by addition of DHT and interestingly, DHT addition also

led to an increase in gelsolin expression. Whether this is an experimental coincidence or due to cell growth is difficult to elucidate based on this data alone.

To further assess the role of gelsolin in androgen mediated wound healing, gelsolin

expression was reduced using gene silencing technology. Multiple controls were put in

place to ensure specificity of gelsolin siRNA. Decreased gelsolin expression was

associated with increased AR expression levels. Given the importance of gelsolin as an

AR co-activator81, it seems possible that the increase in AR expression was

compensatory. To verify this observation, expression of a fellow gelsolin member, Flii

was knocked down in the exact same procedure (data not shown). Interestingly, Flii

ablation was not associated with increase in AR expression. This suggests that the result

may be specific to gelsolin and not another member of the family. It is also noted that

Flii was the only family member tested and not all members of the gelsolin family. As

such, it may still be possible that other members of the family i.e. supervillin, may also

cause an increase in AR expression.

Cells are highly responsive to steroid hormones, which acts as a stimulus for various cellular processes including cell growth. Therefore, adding DHT to HFFs led to an increase in cellular proliferation and migration. To elucidate the importance of gelsolin

100

in AR induced cellular proliferation and migration, gelsolin expression was ablated.

The reduction in gelsolin expression results in significantly less proliferation and

impaired migration. This is in agreement with studies done on gelsolin knockout mice

which have impaired wound healing due to impaired actin depolymerization 78. Given

that gelsolin is also a co-activator of AR, the reduction of gelsolin expression would

have a substantial negative impact on AR activity on cellular proliferation and

migration. Indeed, the addition of DHT to gelsolin knockdown HFFs did not increase

proliferation nor accelerate the rate of wound closure beyond that of normal cells

treated with DHT alone. In other words, DHT was still able to increase proliferation and

accelerate migration but to a lesser degree, suggesting that gelsolin was not essential in

the DHT stimulated cellular proliferation and migration. Although the knockdown of

gelsolin decreased migration, it was probably largely due to its role as an actin

remodeling protein that depolymerizes actin to allow cellular migration.

If gelsolin is a crucial co-activator for AR in wound healing, then the knockdown of

gelsolin expression should arrest cellular proliferation and migration which are the

most important processes during wound healing. However, this was not the case as the

presence of DHT improved cellular proliferation and migration. Taken together, this

study has demonstrated that gelsolin is not directly involved in DHT-stimulated cellular proliferation and migration or that functional redundancy has occurred. This

101

implies that while gelsolin acts as a co-activator for AR, its function can be substituted by other co-activators. Perhaps this may account for the increase in AR expression when gelsolin expression is decreased to compensate for the loss of function. The link between gelsolin and the AR is intriguing. If gelsolin is involved in the androgen mediated wound healing molecular process, then many different biological processes could be affected by manipulation of gelsolin levels in order to control AR activity, which translates into many real-life clinical applications for wound healing.

Due to the outcome of this study showing that gelsolin was not a key re-activator for

AR in a fibroblast wound healing model, the project was not continued. However, future work may involve the analysis using a different approach. Genetically, gelsolin knockout mice could be designed for a wound healing trial with treatment of hormones.

It may also be interesting to determine which AR target genes are activated when gelsolin is the co-activator for that particular expression. This will provide a more definitive answer for the involvement of gelsolin as an AR co-activator.

102

CHAPTER THREE

Flii deficiency improves wound healing and is associated with

higher expression ratio of TGFβ3 to TGFβ1

3.1 Introduction

3.1.1 Flii, a negative regulator of wound healing

Flii is a relatively novel protein involved in a range of cellular processes including its

functions as an actin remodeling protein and nuclear receptor co-activator (reviewed in

180). The full extent of the biological function of Flii is still not fully elucidated. Recent studies using Flii heterozygous null mice, Flii over-expressing transgenic mice, wild-

type (WT) mice in a well established in vivo wound repair model and in vitro research has revealed that Flii is a negative regulator of the wound healing process 110,111. Flii

affects epidermal keratinocyte and dermal fibroblast cellular proliferation and motility

as well as the ability of fibroblasts to produce collagen I. Flii heterozygous mice exhibit

smaller, more contracted wounds with more α-smooth muscle actin positive (αSMA) myofibroblasts, whereas Flii over-expressing mice showed delayed wound closure with increased dermal gape, reduced cellular proliferation and migration 111. Flii deficiency

also correlates with decreased collagen I expression in both in vivo mice wounds as well

103

as in vitro siRNA treated fibroblasts 111. Perhaps the most compelling evidence indicating Flii is detrimental to wound repair is the application of specific Flii neutralizing antibodies to murine incisional wounds. Treatment with Flii neutralizing antibodies improves wound healing, confirming that Flii ablation is beneficial and that manipulation of its activity is possible for therapeutic purposes 111. However, the

underlining mechanisms leading to enhanced healing in Flii heterozygous mice are still

to be determined.

3.1.2 TGFβ and the wound healing process

TGFβ is a powerful multifunctional regulator of cellular growth with effects on

proliferation and differentiation. The most well-known biological function of TGFβ is its

role as a cytokine during the wound healing process, coordinating inflammatory cells as

an initial step to prevent wound infection 181. Even so, the true extent of TGFβ function

is still being unraveled. However, as a cytokine, excessive TGFβ is generally correlated

with increased inflammation. Its influence during the wound healing process is not

only limited to the inflammatory cells but also to epithelial cells such as keratinocytes

and fibroblasts 182-184. TGFβ can affect cellular proliferation and migration, although

various studies have yielded contradictory results indicating that the mechanism of

TGFβ function is still to be completely determined. On the one hand, TGFβ stimulates

the production of integrins (α5β1, αvβ5 and αvβ6) which facilitate the migration of

104

epidermal cells over the provisional wound matrix 185,186. On the other hand, TGFβ

inhibits keratinocyte proliferation in vitro and in vivo 187.

TGFβ is a potential candidate for investigation in the Flii wound healing study for two

main reasons. Firstly, Flii over-expressing mice in the wound healing study exhibit

impaired healing and have increased inflammation. Secondly, 188 reported differential

levels of TGFβ1 in Flii over-expressing wounds, pointing to the involvement of TGFβ in

the wound healing process. As there are three mammalian isoforms of TGFβ, the expression levels of all three isoforms in WT, Flii heterozygous and over-expressing mice wounds will be investigated. It is hypothesized that the impaired healing observed in Flii over-expressing mice is due to excessive expression of TGFβ1.

105

3.2 Materials and Methods

3.2.1 Flii Heterozygous and Over-expressing Mice Generation

The Flii heterozygous (Flii+/-) and over-expressing (FliiTg/+) mice were generated by our

collaborators, Hugh Campbell and Ruth Arkell from Molecular Genetics and

Group for the Molecular Genetics of Development, Research School of Biological

Sciences, Australian National University, Canberra, ACT, Australia 98. WT mice are of

BALB/c genetic background. The methods of generation/breeding are discussed below

in their respective headings.

3.2.2 Flii+/- mice generation

Flii heterozygous mice (Flii+/-) were generated by loss of function mutation in the Flii

gene via homologous recombination in embryonic stem cells and passage of these cells

through the germ line following chimera production. The generation of these mice and

the resulting mutation is described in 98 and a diagram of the targeting strategy is shown in diagram 1A. Animals homozygous for Flii are embryonic lethal and die in utero at embryonic day 7. The strain was therefore maintained by continuous backcross of heterozygous carriers (Flii+/-) to BALB/c mice. All mice used in this project were

heterozygous carriers from backcross generation 10 or later and thus were BALB/c

congenic. The heterozygous mice were identified using three primer PCR sets that

106

1 kb+ ladder mutant allele (538bp)

NOTE: This figure is included on page 107 wild-type allele of the (210bp) print copy of the thesis held in the University of Adelaide Library.

Campbell et al, 2002 98

107

Figure 1

Targeted disruption of the Flii gene. (A) Illustrative representation of the domain structure of the targeting vector, relevant portion of the Flii gene and the targeted allele after homologous recombination. Restriction enzymes sites, BspEI is denoted by B,

EcoRV by E and NcoI by N. Flii exons are represented by the numbered open boxes. The tk-neo and pgk-thymidine kinase casettes are indicated. (B) Three primer PCR indicating wild-type (210kb) and mutant allele (538bp) products. Animals with one wild-type copy of the Flii gene and one mutant allele expressed no more than 50% of the normal wild-type Flii expression levels.

108

amplified products specific to the WT or targeted allele (Diagram 1B). The PCR was performed on DNA extracted from ear biopsies of potential heterozygotes 189. The animals with one wild-type copy of the Flii gene and one mutant copy are presumed to express no more than 50% of the normal Flii gene expression.

3.2.3 FliiTg/+ mice generation

Mice carrying additional copies of the Flii gene were generated by introducing a cosmid construct into the mouse genome via transgenesis. At the time of strain production, the cosmid contained the Flii gene and surrounding sequences with the extent of the construct being defined via restriction mapping. The availability of the mouse genome sequence allowed us to estimate the extent of the cosmid. It is now known that the cosmid contains all of the neighboring SMCR7 gene and parts of the Topo and LLGL1 genes (see Diagram 2). The transgenic strain was backcrossed to BALB/c animals for 10 generations before being intercrossed and homozygous animals were identified via progeny testing. The mouse colony was subsequently maintained by intercross of animals homozygous for the transgene. These animals carry two copies of the mouse

Flii gene and two copies of the human Flii transgene (Flii+/+; FliiTg/Tg). In some cases,

109

NOTE: This figure is included on page 110 of the print copy of the thesis held in

the University of Adelaide Library.

Figure 2

Domain structure of cosmid containing Flii gene which is used to generate Flii transgenic mice. Cosmid contains SMCR7 gene, parts of TOP3A and LLGL1 genes.

Restriction sites on the cosmid are also shown.

Campbell et al, 2002 98

110

these animals were out-crossed to BALB/c animals to generate progeny that carry two

mouse Flii gene and one copy of human Flii transgene (Flii+/+; FliiTg/+ are denoted as

FliiTg/+ throughout the thesis).

3.2.4 Murine Incisional Wound Surgery

All animal experiments were approved by the Adelaide Women’s and Children’s

Hospital Animal Care and Ethics Committee and the Australian National University

Animal Ethics Committee following the Australian Code of Practice for the Care and

Use of Animals for Scientific Purposes. The mice were housed in the Women’s and

Children’s Hospital animal house and cages were cleaned and food and water replaced

daily.

Surgery was performed on 72 twelve weeks old mice, 24 wild-type (WT), 24 Flii

heterozygous (Flii+/-) and 24 transgenic (FliiTg/+) mice. Four timepoints, 3, 7, 14 and 21 days post-wounding were chosen to represent the different periods in wound healing.

n=6 per timepoint, 2 wounds created per mice. Mice were weighed upon arrival, prior

to surgery and post-surgery at timepoints stated above to monitor their well-being.

Mouse behavior was also observed regularly. Before surgery, the mice were induced

with anesthesia using 5% isofluorane at 2L oxygen per minute. Anesthesia was

maintained using 2% isofluorane at 500ml oxygen per minute throughout surgery. The

111

mice were then shaved. Two equidistant, 1cm full thickness incisions were made

through the skin to the panniculus carnosus on the flanks of the mice extending 3-4 cm from the base of the skull, 1cm either side of the spinal column and left to heal by secondary intention (i.e. wounds left unopposed and not sutured). At designated day of

sacrifice, the mice were weighed and euthanized using CO2 and cervical dislocation.

The mice were then re-shaved and photographs of the wounds were taken prior to

excision. Both left and right wounds were excised and bisected, half was used for

histological analyses and the other half was snap frozen in liquid nitrogen and stored at

-80°C for biochemical analyses. For each mouse group, unwounded skin was collected

from the WT, Flii+/- and FliiTg/+ mice and divided as stated above.

3.2.5 Histological Processing

Tissue collected from surgery was fixed in 10% formalin and processed in a Leica

TP1020 tissue processor which dehydrated the tissues in graded alcohol series (70% for

120mins, 80% for 60mins, 90% for 105mins and 100% for 240mins), cleared in transitional solvent xylene for 180mins followed by 240mins of tissue infiltration with paraffin wax. Tissue sections (4µm) were cut from paraffin-embedded fixed tissue using

Leica RM2235 microtome. Prior to staining, skin sections were dewaxed by a series of xylene (30mins), and graduated ethanol washes (100%, 1min, 70%, 1 min & 30%, 1 min)

112

before further processing. Skin sections were either stained with haemotoxylin and

eosin (H&E) or with antibodies listed in Table 3.1.

3.2.6 Immunohistochemistry

Immunohistochemical staining of TGFβ1, TGFβ2, TGFβ3, TβRI and TβRII expression

was performed on all WT, Flii+/- and FliiTg/+ mice skin at day 0, day 3, day 7, day 14 and

day 21 post-wounding and analysis of fluorescence assessed in the dermis of both

wounded and unwounded skin. Antigen retrieval was required for these antigens.

Sections were dewaxed by a series of xylene (30mins), and graduated ethanol washes

(100% for 1min, 70% for 1 min & 30% for 1 min) before rinsing in 1x PBS and pre-treated

with 250ml Target Retrival Solution (TRS) (2.8g Citric Acid, 3.76g Glycine, 0.372g

EDTA, pH 5.9 in 1L 1x PBS) solution. The sections were then microwaved on high

setting for 2mins after which a “ballast” pot of water was added to help absorb some

heat and to prevent damage to sections and pretreatment continued for 2 x 5mins in

microwave with regular “airing” to prevent overheating. Temperature was maintained

at 94°C. Sections were cooled to 50°C on ice before rinsing in 1x PBS and enzymatic

digestion for 3mins with 0.0625g of Trypsin (Sigma Aldrich, Sydney, Australia)

dissolved in pre-warmed 1x PBS at 37°C. Following digestion, sections were washed in

1x PBS and incubated with normal horse serum (NHS, 3% dissolved in 1x PBS) blocking

solution

113

Concentration used Raised Antibody Manufacturer Catalog # in Western Immunostaining Santa Cruz TGFβ1 sc-146 Rabbit 1µg/ml 2µg/ml Biotechnology

TGFβ2 R&D Systems AB-12-NA Rabbit 1µg/ml 2µg/ml

AB-244- TGFβ3 R&D Systems Goat 1µg/ml 2µg/ml NA Santa Cruz TGFβ Receptor I sc-399 Rabbit 1µg/ml 2µg/ml Biotechnology Santa Cruz TGFβ Receptor II sc-1700 Rabbit 1µg/ml 2µg/ml Biotechnology 2° Anti-rabbit Invitrogen A11008 Goat N/A 2µg/ml Alexa Fluor 488 Biotinylated Vector 2° Anti-mouse BA-2000 Horse N/A 2µg/ml Laboratories IgG Strepdavidin Invitrogen S32355 N/A N/A 2µg/ml Alexa Fluor 555

Table 3.1

Information of antibodies used in western analyses and immunofluorescent stainings.

114

for 30 mins at room temperature. Refer to Table 3.1 for antibody information. This was

followed by another wash in 1x PBS before incubation with primary antibodies (3%

NHS dissolved in 1x PBS) in a moist box overnight at 4°C. Sections were then washed 3

x 2mins in 1x PBS followed by incubation with biotinylated secondary antibodies (1x

PBS) for 1hr at room temperature. Sections were then washed again 3 x 2mins in 1x PBS

and incubated with cy3 Streptavidin (Sigma Aldrich, Sydney, Australia) conjugate (1x

PBS) for 45mins at room temperature. This was followed by final 3 x 2mins washes before mounting sections using Dako Fluorescent Mounting Medium (DAKO, Botany,

Australia). Slides were then stored in dark at -20°C. Integrated fluorescence intensity was determined using AnalySIS software (Soft-Imaging System GmbH, Munster,

Germany). Negative controls were included to demonstrate antibody staining specificity. Control samples undergo the exact staining procedure omitting either the primary or the secondary antibody. All control samples had negligible immunofluorescence.

3.2.7 Cell culture

Human Foreskin Fibroblasts (HFFs) were cultured in Dulbecco’s modified Eagle’s

Medium (DMEM) supplemented with 10% fetal calf serum (FCS) and antibiotics (100U

penicillin and 100ug/500ml streptomycin). Cell cultures were incubated at 37°C and 5%

115

CO2. Cells were serum starved in DMEM containing antibiotics for at least 3 hours or

otherwise stated prior to start of experimenting to synchronize cells.

3.2.8 siRNA knockdown

HFFs were seeded into 6 well tissue culture plates and cultured overnight to achieve

30% to 50% confluence at time of transfection. Sequence of Flii siRNA are as follows;

forward: 5’-GCU GGA ACA CUU GUC UGU GTT-3’, reverse: 5’-CAC AGA CAA GUG

UUC CAG CTT-3’ 105. siRNA were transfected into the cells using Lipofectamine 2000

(Invitrogen, Carlsbad, USA). Both siRNA and Lipofectamine 2000 were diluted in Opti-

MEM I Reduced Serum Medium (Invitrogen, Carlsbad, USA). 250µl of siRNA

(optimized to 100nM per well) was mixed with 4µg of Lipofectamine 2000 diluted in

250µl Opti-MEM and were allowed to complex at room temperature for 20 minutes.

500µl of siRNA:Lipofectamine 2000 complex was then added to each well, mixed and

cells incubated for 6 hours before replacing transfection media with DMEM containing

10% FCS only. Cells were incubated for 48 hours prior to gene knockdown assessment.

3.2.9 RNA extraction

RNA was extracted from Human Foreskin Fibroblasts. RNA extraction from cell

cultures only required scraping cells from culture flasks after adding Trizol Reagent

116

(Invitrogen, Victoria, Australia). Samples were transferred into fresh eppendorf tubes

and centrifuged at 12,000g at 4°C for 10mins to remove cell debris. The samples were

incubated for 5mins at room temperature before adding 200µl of chloroform to each

tube and were mixed thoroughly by hand for 15secs. The samples were kept at room

temperature for 3mins and centrifuged at 12,000g for 15mins at 4°C. The aqueous phase

containing RNA was transferred into a fresh tube and 500µl of isopropanol added to

precipitate the RNA. The samples were then incubated at room temperature for 10mins

before centrifuging at 12,000g for 10mins at 4°C. The supernatant was discarded and the

residual pellet washed with 1ml of 75% ethanol. Finally, samples were centrifuged at

7500g for 5mins at 4°C and supernatant discarded. The pellet was dried and re- dissolved in 50µl DEPC water.

3.2.10 DNase Treatment and RNA Quantitation

RNA samples obtained were subjected to DNA-free DNase Treatment and Removal Kit

(Ambion, TX, USA) as instructed by the manufacturer to remove any contaminating

genomic DNA. Firstly, RNA samples were treated with 0.1 volume of 10x DNase Buffer

and 1µl of rDNase I and incubated at 37°C for 30mins. Following this, 0.1 volume of

DNAse Inactivating Reagent was added to the samples and incubated at 2mins at room

temperature with occasional mixing. The samples were then centrifuged at 10,000g for

90secs and supernatant transferred to fresh tubes. RNA was quantitated by diluting 1 in

117

20 with RNase free water and 100µl duplicates were quantified using a Pharmacia

Biotech GeneQuant RNA/DNA Calculator using RNase-free water as a blank.

Absorbance at 260nm and 280nm were measured that quantify RNA absorbance as

µg/µl concentration. Purity of RNA was confirmed by the A260/A280 ratio and a value

between 1.7 to 2.0 indicates good RNA quality.

3.2.11 Complementary Deoxyribonucleic Acid (cDNA) Synthesis

cDNA was synthesized from RNA using reverse transcription. Each reaction contains

1µg of RNA with 4µl 2.5µM dNTPs (dATP, dCTP, dGTP and dTTP, 100mM each,

Promega, WI, USA) and 2µl Oligo(dt)12-18 Primer (25µg at 0.5µg/µl, Invitrogen, Victoria,

Australia). This was heated at 85°C for 3mins and placed in ice immediately. 2µl 10x

Stratascript Buffer (Stratagene, Epson, UK), 1µl RNasin (Promega, WI, USA) and 1µl

Stratascript Reverse Transcriptase (Stratagene, Epson, UK) were added to the mixture

and heated at 42°C for 60mins followed by 92°C for 10mins before cooling it down on

ice. A control sample was prepared with the reagents above with the exclusion of

reverse transcriptase for use in the Real-Time quantitative-Polymerase Chain Reaction

(RTq-PCR) as a negative control.

118

3.2.12 Real-Time quantitative-Polymerase Chain Reaction (RTq-PCR)

Each PCR reaction tube containing cDNA was set up to a final concentration of 1x SYBR

Green, 1x Amplitag PCR buffer, 3mM MgCl2, 5mM dNTPs, 0.9µM primers (forward

and reverse), 1.25 Units of AmpliTag Gold DNA polymerase in 25µl of H2O. Primer

sequences used in this chapter is shown in Table 3.2. The cycle conditions are as follows;

an initial denaturation at 95°C for 15 minutes, 35 cycles of 95°C for 25 seconds, 60°C for

30 seconds, 72°C for 30 seconds and at the final cycle, an additional 5 minutes at 72°C before a melt from 72°C to 99°C at 30 seconds at each degree step.

3.2.13 Statistical Analysis

All statistical differences were determined using the Student’s t-test or ANOVA. P

value of less than 0.05 was considered significant.

119

Forward or Primer Sequence 5' - 3' Reverse forward CCT CCT ACA GCT AGC AGG TTA TCA AC (Flii) reverse GCA TGT GCT GGA TAT ATA CCT GGC AG forward CAG ACA GCC CCT GCC AGC ACC C Gelsolin reverse GAG TTC AGT GCA CCA GCC TTA GGC forward GGT TGG ATG GCA AGC ATG TG Cyclophillin A reverse TGC TGG TCT TGC CAT TCC TG forward CTA CGA GGC GTC ATC CTC CCG c-Fos reverse AGC TCC CTC CGG TTG CGG CAT forward GAA ACG ACC TTC TAT GAC GAT GCC CTC AA c-Jun reverse GAA CCC CTC CTG CTC ATC TGT CAC GTT CTT forward GTT GGA CGA GCT GGA GAA GG Smad 3 reverse TGC TGT GGT TCA TCT GGT GG forward ACG GCC ATC TTC AGC ACC AC Smad 4 reverse AGA ATG CAC AAT CGC CGG AG forward GCT CAC GCA CTC GGT GCT CA Smad 7 reverse CCA GGC TCC AGA AGA AGT TG

Table 3.2

Primer sequences used in RT-qPCR

120

3.3 Results

3.3.1 Time-course analyses of TGFβ1 expression in wounded mice skin

Murine skin sections were stained with haemotoxylin and eosin to illustrate the key differences of unwounded and wounded skin sections (Figure 3.1A, C). A typical skin section is represented by three distinct components. The epidermis consists of producing stratified keratinocytes which overlies the dermis. This middle layer is composed of loosely arranged connective tissues, fibroblasts, collagens, nerves, blood vessels (Figure 3.1A, B, arrowheads) and other adnexal structures such as hair follicles

(Figure 3.1A, arrows). Immune cells such as macrophages and mast cells are also found in the dermis. Below the dermis is the subcutaneous layer which is primarily composed of adipose tissue. This is seen as the layer between the dermis and the panniculus muscle as shown in Figure 3.1A. In a wounded skin section, the scab tissue and a thick re-epithelializing epidermis can be clearly seen as shown in Figure 3.1C. No hair follicle could be seen at the wound site due to the complete removal in a full thickness injury.

TGFβ1 staining was done to show the relative position of an immunofluorescent image on a skin section (Figure 3.1,B, D). All immunofluorescent images are taken at or adjacent to the position defined by the box in Figure 3.1A,C.

121

ep ep d

d

p A B

C ep ep s wm

p wm D

122

Figure 3.1

Histology and immunohistochemistry of murine skin sections. Skin biopsies from WT mice were fixed in formalin, embedded in paraffin wax, sectioned (4µm) and stained either with hematoxylin and eosin (H&E) to illustrate normal and wounded tissue morphology, or anti-TGFβ1 antibodies to show the relative position of an immunofluorescent image on a skin section. (A,C) Typical H&E staining of a normal

(unwounded) and wounded murine skin section. (B,D) Immunohistochemical staining of TGFβ1 in a relative position as indicated by a box in (A) and (C) respectively. In all images, ep indicates position of epidermis, d indicates position of dermis, wm indicates position of wound matrix, p indicates the position of the panniculus and s indicates the position of the scab tissue. Arrowheads indicate blood vessels and arrows indicate hair follicles in (A). Dotted white or black line separates the epidermis from the dermis or the wound matrix in all images. Scale bar = 50 µm in (A, C), (B, D).

123

negligible in WT, Flii+/- and FliiTg/+ unwounded skin (day 0) indicating low TGFβ1

expression levels (Figure 3.2A-C). As expected, TGFβ1 expression significantly

increased upon wounding. The expression of TGFβ1 was highest at day 7 post-

wounding in Flii+/-, WT and FliiTg/+ wounds and is shown in Figure 3.2D-F. At day 7,

TGFβ1 expression was observed to be mainly localized to the provisional wound matrix

in the dermis where the major TGFβ secreting cells such as the fibroblasts reside. The

increase in TGFβ1 expression remain elevated in the provisional wound matrix until 14

days post-wounding before returning to normal basal expression at 21 days post-

wounding (Figure 3.2G). It was found that TGFβ1 expression was significantly elevated

particularly in FliiTg/+ wounds particularly at day 7 post-wounding and remained

elevated at day 14 and day 21 although reduced from its peak at day 7. It was noted that

Flii+/- wounds expressed significantly lower levels of TGFβ1 levels at day 7 compared to

WT and FliiTg/+ wounds throughout the wound healing time-course.

3.3.2 Time-course analyses of TGFβ2 expression in wounded mice skin

TGFβ2 expression was negligible in WT, Flii+/- and FliiTg/+ unwounded skin (day 0)

(Figure 3.3A-C). At day 7, TGFβ2 expression was significantly elevated in all WT, Flii+/-

and FliiTg/+ wounds compared to unwounded skin and is shown in Figure 3.3D-F. In day

7 wounds, the majority of TGFβ2 expression was localized to the provisional wound

124

Flii+/- WT FliiTg/+

A ep B C ep ep Day 0 d d d

D E ep F ep

Day 7 wm wm wm

G Flii+/- 100 *c-e WT

80 FliiTg/+ *f-g 60 *h-i Fluorescence *a-b

1 40 Intensity/unit area Intensity/unit TGF β 20

0 0 3 7 14 21 Days

125

Figure 3.2

Immunofluorescence staining of TGFβ1 in WT, Flii+/- and FliiTg/+ unwounded and wounded skin sections. (A-C) Unwounded skin sections showing negligible TGFβ1 staining. (D-F) Day 7 immunofluorescence of TGFβ1 in WT, Flii+/- and FliiTg/+ wounds.

(G) Quantitated immunofluorescent staining intensity of TGFβ1 in WT, Flii+/- and FliiTg/+ wounds at day 0 (unwounded), 3, 7, 14 and 21 post-wounding. In all images, ep indicates position of epidermis, d indicates position of dermis and wm indicates position of wound matrix. Dotted white or black line separates the epidermis from the dermis or the wound matrix in all images. Results represent mean ± S.E.M. n = 6 for each group per time-point. *a = WT vs FliiTg/+, *b = Flii+/- vs FliiTg/+, c = WT vs FliiTg/+, d =

WT vs Flii+/-, e = Flii+/- vs FliiTg/+, f = WT vs FliiTg/+, g = Flii+/- vs FliiTg/+, h = WT vs FliiTg/+, i =

Flii+/- vs FliiTg/+.*p < 0.05. Scale bar = 50 µm in (F).

126

Flii+/- WT FliiTg/+ A ep B ep C ep

d d d Day 0

D E ep F ep ep

wm wm Day 7 wm

G *a-c 100 Flii+/- *d-e WT 80 FliiTg/+

60

Fluorescence 40 2 Intensity/unit area Intensity/unit TGF β 20

0 0 3 7 14 21 Days

127

Figure 3.3

Immunofluorescence staining of TGFβ2 in WT, Flii+/- and FliiTg/+ unwounded and wounded skin sections. (A-C) Unwounded skin sections showing negligible TGFβ2 staining. (D-F) Day 7 immunofluorescence of TGFβ2 in WT, Flii+/- and FliiTg/+ wounds.

(G) Quantitated immunofluorescent staining intensity of TGFβ2 in WT, Flii+/- and FliiTg/+ wounds at day 0 (unwounded), 3, 7, 14 and 21 post-wounding. In all images, ep indicates position of epidermis, d indicates position of dermis and wm indicates position of wound matrix. Dotted white or black line separates the epidermis from the dermis or the wound matrix in all images. Results represent mean ± S.E.M. n = 6 for each group per time-point. *a = WT vs FliiTg/+, *b = WT vs Flii+/-, *c = Flii+/- vs FliiTg/+, d =

WT vs FliiTg/+, e = Flii+/- vs FliiTg/+. *p < 0.05. Scale bar = 50µm in (F).

128

matrix as well as the surrounding dermis. During the wound healing time-course experiment spanning 21 days, TGFβ2 expression was significantly increased at day 3 post-wounding and was also maintained at higher levels throughout the time-course

(Figure 3.3G). Unlike TGFβ1, TGFβ2 levels remained significantly elevated at day 21 in all groups. TGFβ2 expression increased up to day 7 and this was particularly striking in

FliiTg/+ wounds which was significantly higher than WT and Flii+/- respectively. TGFβ2

expression reduced at day 14 and 21 before returning to baseline levels.

3.3.3 Time-course analyses of TGFβ3 expression in wounded mice skin

Similar to TGFβ1 and TGFβ2, TGFβ3 immunofluorescence was minimal in WT, Flii+/-

and FliiTg/+ unwounded skin (day 0) (Figure 3.4A-C). Representative images of day 7

wounds are shown in Figure 3.4D-F. Staining for TGFβ3 is similar to TGFβ1 and TGFβ2

staining where TGFβ3 expression increased in response to wounding. The majority of

TGFβ3 expression was localized to the provisional wound matrix and the surrounding

dermis. Analysis of the expression profile of TGFβ3 during wound healing showed a

significant increase in expression of this isoform in day 7 wounds until day 14 before

returning to basal levels at day 21 post-wounding (Figure 3.4G). Both WT and FliiTg/+

wounds also expressed higher TGFβ3 at day 3, 7 and 14 but returned to basal levels at

day 21. WT and FliiTg/+ wounds had no noticeable difference throughout the wound

healing time-course. However, in contrast to what was observed with TGFβ1 and

129

Flii+/- WT FliiTg/+

A ep B ep C ep

Day 0 d d d

ep D ep E ep F

wm Day 7 wm wm

G Flii+/- 100 WT *a-b *c-d 80 FliiTg/+

60 Fluorescence

3 40

Intensity/unit area Intensity/unit TGF β 20

0 0 3 7 14 21

Days

130

Figure 3.4

Immunofluorescence staining of TGFβ3 in WT, Flii+/- and FliiTg/+ unwounded and wounded skin sections. (A-C) Unwounded skin sections showing negligible TGFβ3 staining. (D-F) Day 7 immunofluorescence of TGFβ3 in WT, Flii+/- and FliiTg/+ wounds.

(G) Quantitated immunofluorescent staining intensity of TGFβ3 in WT, Flii+/- and FliiTg/+ wounds at day 0 (unwounded), 3, 7, 14 and 21 post-wounding. In all images, ep indicates position of epidermis, d indicates position of dermis and wm indicates position of wound matrix. Dotted white or black line separates the epidermis from the dermis or the wound matrix in all images. Results represent mean ± S.E.M. n = 6 for each group per time-point. *a = WT vs FliiTg/+,*b = Flii+/- vs FliiTg/+, c = WT vs FliiTg/+, d =

Flii+/- vs FliiTg/+.*p < 0.05. Scale bar = 50µm in (F).

131

TGFβ2 isoforms, Flii+/- wounds had significantly higher levels of TGFβ3 at day 7 and 14

post-wounding than WT and FliiTg/+. These levels of TGFβ3 were more than doubled at day 7 compared to the unwounded samples.

In our immunofluorescent staining results, we found that day 7 exhibited the most differences between TGFβ1, TGFβ2 and TGFβ3. Therefore, the staining intensities of each TGFβ isoform were compared between one another for each individual WT, Flii+/-

and FliiTg/+ wounds at day 7 post-wounding. Comparative expression of TGFβ1, TGFβ2

and TGFβ3 in WT, Flii+/- and FliiTg/+ wounds at day 7 are shown in Figure 3.5. Flii+/-

wounds at day 7 expressed lower levels of TGFβ1 and higher levels of TGFβ3 compared

to WT and FliiTg/+ wounds. TGFβ2 expression was also marginally lower than WT and

FliiTg/+ wounds at day 7. Conversely, FliiTg/+ wounds had an entirely different TGFβ

expression profile to the Flii+/- wounds. FliiTg/+ wounds had higher levels of TGFβ1

compared to WT and Flii+/- wounds. TGFβ2 expression was also slightly elevated in

FliiTg/+ wounds. Although TGFβ3 levels were slightly higher than in WT wounds, it was

significantly lower than that of Flii+/- wounds. Therefore, it can be clearly observed that

Flii+/- wounds had a low TGFβ1 to high TGFβ3 expression ratio which was opposite to

FliiTg/+ wounds with an expression ratio of high TGFβ1 to low TGFβ3.

132

Day 7

200 *e *c-d

150 *f TGF-β1 100 *a-b TGF-β2

50 TGF-β3

0 Fluorescence Intensity (% of Control) Flii+/- WT FliiTg/+

Figure 3.5

Fluorescent intensity comparison of TGFβ1, TGFβ2 and TGFβ3 expressions between

WT, Flii+/- and FliiTg/+ sections in day 7 post-wounded skin and expressed as a

percentage of control. Flii+/- wounded skin expressed a lower amount of TGFβ1 and

considerably higher TGFβ3 whereas FliiTg/+ wounded skin expressed higher amount of

TGFβ1 and lower levels of TGFβ3. Results represent mean ± S.E.M. n = 6 for each group

per time-point. *a = Flii+/- vs WT, *b = Flii+/- vs FliiTg/+, *c = Flii+/- vs WT, *d = Flii+/- vs

FliiTg/+, *e = FliiTg/+ vs WT, *f = FliiTg/+ vs WT. *p < 0.05.

133

3.3.4 Time-course analyses of TβRI and TβRII in wounded mice skin

TβRI and TβRII are the main receptors responsible for transducing TGFβ signals and

were also investigated in this study. Immunohistochemical staining of TβRI and TβRII

in unwounded skin showed that both TβRI and TβRII were localized to the epidermis,

endothelial cells lining the blood vessels and hair follicles (Figure 3.6 A-C & Figure

3.7A-C). Minimal staining were observed in the dermis of unwounded skin. In addition,

no significant difference in TβRI and TβRII expression was observed between WT, Flii+/-

and FliiTg/+ unwounded skins. TβRI and TβRII expression was increased in response to

wounding. In wounded skin sections, staining for TβRI and TβRII was observed to be

localized to the newly formed epidermis as well as the underlining provisional wound

matrix.

By analyzing the TβRI expression profile, it was observed that TβRI expression

increased slightly at day 3 and become significantly higher at day 7 before decreasing at

day 14 and 21 post-wounding in all WT, Flii+/- and FliiTg/+ wounds (Figure 3.6G). There were no differences in TβRI expression between WT, Flii+/- and FliiTg/+ wounds at all

time-points. It was also noted that at day 21 post-wounding, TβRI remained slightly elevated compared to the basal expression levels in unwounded skin. TβRII also had a similar expression profile (Figure 3.7G). In wounded skin, TβRII expression was

134

Flii+/- WT FliiTg/+

A ep B ep C ep

Day 0 d d d

D ep E ep F ep

Day 7 wm wm wm

G Flii+/- 100 WT

80 FliiTg/+

60

Fluorescence 40 RI β T Intensity/unit area Intensity/unit 20

0 0 3 7 14 21 Days

135

Figure 3.6

Immunofluorescence staining of TβRI in WT, Flii+/- and FliiTg/+ unwounded and wounded skin sections. (A-C) Unwounded skin sections showing negligible TβRI staining. (D-F) Day 7 immunofluorescence of TβRI in WT, Flii+/- and FliiTg/+ wounds. (G)

Quantitated immunofluorescent staining intensity of TβRI in WT, Flii+/- and FliiTg/+ wounds at day 0 (unwounded), 3, 7, 14 and 21 post-wounding. In all images, ep indicates position of epidermis, d indicates position of dermis and wm indicates position of wound matrix. Dotted white or black line separates the epidermis from the dermis or the wound matrix in all images. Results represent mean ± S.E.M. n = 6 for each group per time-point. Scale bar = 50µm in (F).

136

Flii+/- WT FliiTg/+

A ep B ep C ep

d Day 0 d d

D ep E ep F ep

Day 7 wm wm wm

G Flii+/- 100 WT

80 FliiTg/+

60

40 RII Fluorescence RII β T Intensity/unit area Intensity/unit 20

0 01 23 37 144 215 Days

137

Figure 3.7

Immunofluorescence staining of TβRII in WT, Flii+/- and FliiTg/+ unwounded and

wounded skin sections. (A-C) Unwounded skin sections showing negligible TβRII

staining. (D-F) Day 7 immunofluorescence of TβRII in WT, Flii+/- and FliiTg/+ wounds. (G)

Quantitated immunofluorescent staining intensity of TβRII in WT, Flii+/- and FliiTg/+

wounds at day 0 (unwounded), 3, 7, 14 and 21 post-wounding. In all images, ep

indicates position of epidermis, d indicates position of dermis and wm indicates

position of wound matrix. Dotted white or black line separates the epidermis from the

dermis or the wound matrix in all images. Results represent mean ± S.E.M. n = 6 for

each group per time-point. Scale bar = 50µm in (F).

138

increased at day 3 and expression peaked at day 7 post-wounding before declining at day 14 and reaching basal levels at day 21. There were also no noticeable differences between WT, Flii+/- and FliiTg/+ wounds at all time-points.

3.3.5 Flii knockdown alters TGFβ gene expression.

To further substantiate that manipulation of Flii affects TGFβ expression, an in vitro Flii

siRNA gene knockdown method was employed. Flii gene expression in normal human

foreskin fibroblasts (HFFs) was knocked down and mRNA quantified using real-time

quantitative polymerase chain reaction (RT-qPCR). Flii gene expression levels after

knockdown were approximately 22%, which was significantly lower than its normal

levels (Figure 3.8A). A scrambled siRNA sequence was used as a control to show that

sham siRNA knockdown process had no effect on Flii gene expression. Gelsolin gene

expression was also included as a specificity control and showed no difference (Figure

3.8B). However, changes in TGFβ isoform mRNA expressions in cells with reduced Flii

gene expression were clearly observed (Figure 3.8B). Both TGFβ1 and TGFβ2 levels

were significantly decreased and mRNA expression levels were approximately at 56%

and 62% of original expression. On the contrary, TGFβ3 mRNA levels were significantly

increased (37%) in response to Flii gene knockdown. This complemented the in vivo

Flii+/- wound healing results where mice heterozygous for Flii gene expression had higher TGFβ3 protein compared to TGFβ1 and TGFβ2 expression.

139

A 120

100

80

60 * 40

Fold Change (% control) to Change Fold 20

0 Flii siRNA Scrambled siRNA

B 160 * 140

120

100

* 80 * 60

40 * Fold Change (% Control) to Change Fold 20

0 Flii Gelsolin TGF-β1 TGF-β2 TGF-β3

140

Figure 3.8

Effects of Flii siRNA knockdown on TGFβ and its associated genes. Flii gene expression was knocked down using siRNA in normal human foreskin fibroblasts. Gene expressions of TGFβs were quantitated using RTq-PCR. Results are expressed as a percentage to control which is represented by the dotted line at 100%. (A) Level of Flii gene knockdown using siRNA. Scrambled siRNA sequence is included as a control to show that process of siRNA knockdown does not decrease Flii gene expression. (B) mRNA expression levels of TGFβ1, TGFβ2 and TGFβ3 in Flii knockdown HFFs.

Gelsolin, a member of the family is included as a control gene to show target specificity of Flii siRNA. Results represent mean ± S.E.M. n = 6 for each group, *p < 0.05.

141

3.4 Discussion

The actin remodeling protein Flii is a negative regulator of wound healing 111. Healing was enhanced in mice heterozygous for Flii gene whereas mice over-expressing Flii had larger wounds due to impaired healing. It is understood that difference in cellular characteristics such as migration and proliferation contributed to the above outcome.

However, the molecular mechanism for such a variation in cellular characteristics remains unknown. In a separate study by 188, higher levels of TGFβ1 were found in mice

over-expressing Flii which provided a clue, suggesting that TGFβ may be involved in

the Flii mediated wound healing process. As a result, it was hypothesized that the

enhanced healing in mice heterozygous for Flii gene is due to the levels of TGFβ1

present during wound healing. This study has provided more insight into the enhanced

healing in Flii heterozygous mice by profiling expression of TGFβ isoforms and their

receptors TβRI and TβRII using an in vivo time-course wound healing trial.

During the trial, TGFβ1 expression was dramatically increased in mice over-expressing

Flii, consistent with results from 188. However, increases were also observed in WT and

Flii heterozygous mice, albeit in lesser amounts than Flii over-expressing mice. This

observation was also similar for TGFβ2 expression. Under normal wound healing

conditions, TGFβ1 is upregulated in response to wounding. Indeed, reduced TGFβ1 can

142

lead to weaker wound strength and impaired healing 121. However, this is more likely due to the multifocal inflammation nature of TGFβ1 mice as no differences in healing between wild-type and TGFβ1 knockout mice occur until 10 days later when inflammation sets in, consequently affecting wound healing 190,191. Conversely, excessive

TGFβ1 has been shown to directly promote scar formation 117,192. It has also been reported that the reduction of TGFβ1 expression led to an improvement in the healing of cutaneous wounds due to a lower inflammatory response 193. Therefore, this may explain the impaired healing outcome of Flii over-expressing mice wounds which express more TGFβ1, or the lack thereof in Flii heterozygous mice wounds, which exhibited improved healing.

TGFβ1 only provides part of the story; expression of TGFβ2 and TGFβ3 also have to be considered. The function of TGFβ2 remains less well understood compared to TGFβ1 and TGFβ3 in the process of wound healing. Although, it is a less potent cytokine compared to both TGFβ1 and TGFβ3, it still contributes to the acceleration of scar formation and increased collagen I production 194. For this reason, it too is not desirable in high levels during wound healing. Similarly to the above TGFβ1 results, expression is highest in mice over-expressing Flii and lowest in Flii heterozygous mice. Therefore, higher TGFβ2 expression may also contribute or exacerbate the effects of TGFβ1 in Flii over-expressing mice wounds to result in impaired healing. The higher levels of both

143

TGFβ1 and TGFβ2, as seen in Flii over-expressing mice wounds, have been shown to

promote scarring 117. The authors used neutralizing antibodies to TGFβ1 and TGFβ2 and

showed that by eliminating the activities of these cytokines, they could reduce the

detrimental effects of TGFβ1 and TGFβ2 during wound healing.

Expression of TGFβ3 in Flii heterozygous and over-expressing mice provided an

interesting and novel result during the wound healing trial. Various studies have

revealed the beneficial properties of TGFβ3 in enhancing wound healing as well as

improving the outcome of scar formation 116,118,195,196. The mechanism underlying TGFβ3s

beneficial properties is still not clear. However, one interesting point is that TGFβ1 and

TGFβ2 are more associated with adult wound healing and TGFβ3 is more closely associated with fetal scarless wound healing 115. Therefore, these findings may provide

an explanation for the improvement in Flii heterozygous mice wounds in which this

study showed higher levels of TGFβ3 than WT and Flii over-expressing mice wounds.

To further investigate TGFβ signaling responses, the expression profile of its receptors,

TβRI and TβRII were also investigated. Although both receptors are important in

determining the outcome of wound healing by altering downstream signaling activities

of TGFβs, no significant differences between the mice groups were identified. This

suggests that TβRI and TβRII expressions is not directly affected by the expression of

144

Flii. The differences in wound healing observed in WT, Flii heterozygous and over-

expressing mice are more likely to be attributable to the TGFβs.

A direct comparison of all three TGF isoforms were performed in WT, Flii heterozygous

and over-expressing mice on day 7 of the wound healing time-course. Day 7 was chosen

as a comparison time-point due as the period coincides with the proliferative and

remodeling phase which is essentially the most important part of the wound healing

process. Coincidentally, it is also the time-point that showed the most differences

between TGFβ expressions in each mice group. TGFβ1 and TGFβ2 levels were relatively

low in Flii heterozygous mice wounds but a high expression of TGFβ3 was observed.

Flii over-expressing mice wounds on the other hand had significantly higher TGFβ1

expression but lower TGFβ3 expression compared to Flii+/- wounds. These low TGFβ1

and TGFβ2 to high TGFβ3 levels are consistent with the in vitro Flii siRNA knockdown

results, which also showed lower levels of TGFβ1 and TGFβ2 to higher levels of TGFβ3.

Given this, it is most probable that the ratio of TGFβ1/TGFβ2 to TGFβ3 contributes to

the wound healing outcome observed in Flii+/- mice.

TGFβ is a double edge sword in the wound healing process which is required for the coordination of inflammatory cells, yet high levels of TGFβs cause excessive inflammation which contributes to undesirable scarring, a consequence of any adult

145

wound healing process. This study now reveals that the relationship between Flii and

TGFβ expression and the effects on the outcome of wound healing in WT, Flii heterozygous and over-expressing mice.

146

CHAPTER FOUR

Flii modulation of TGFβ expression is achieved by direct

association with proteins involved in the expression and

activity of TGFβ.

4.1 Introduction

4.1.1 Binding properties of Flii protein

The Flii protein consists of two distinct domains, an N-terminal LRR domain and a C- terminal gelsolin like domain which binds actin 93,103. The LRR domain is involved in protein recognition and is found in proteins with diverse cellular functions 197. The LRR domain therefore provides Flii with functions distinct from those of an actin remodeling protein. The capabilities of Flii as an interacting protein have been investigated and the results have implicated Flii in various cellular processes 180. Interestingly, many of these processes are involved in signal transduction and are regulated by direct interactions with the Flii protein. The first of these processes is the role of Flii as a co-activator in nuclear receptor signaling with estrogen and thyroid hormone receptor as well as transcription factors, CARM1 and GRIP1 have been identified as Flii binding partners

105. In addition, proinflammatory caspase-1 and caspase-11 have also been shown to

147

interact with Flii 108. This finding suggests a role for Flii in the immune response.

Furthermore, Flii was also found to interact with MyD88 which is a modulator for Toll-

like Receptors (TLRs) involved in innate immune system 198. The interaction of Flii with

proteins involved in the cell cycle has also been reported. Using epitope-tag tandem mass spectrometry, Seward et al (2008) identified Flii as a binding partner of CaMK-II, a

protein kinase involved in cell cycle progression. Perhaps the most exciting of

interactions is the identification of Flii as a substrate of a cytokine-independent survival

kinase (CISK), a downstream target of phosphoinositol 3-kinase 107. This is the first

description of Flii as a substrate that can be post-translationally modified which further

implicates Flii involvement as a signaling molecule. While these studies did not specify

the between Flii and other proteins, it is most likely mediated through the

LRR domain due to its protein recognition properties.

4.1.2 Properties of the LRR motif

The uniqueness of Flii in the gelsolin family of actin binding proteins is due to the

presence of a 16 tandem 23 amino acid LRR motif. Proteins containing LRR motifs have

diverse cellular localizations including extracellular, cytoplasmic, transmembrane and

nuclear localization. This is in addition to the extraordinary range of cellular functions

where LRR containing proteins are involved which includes receptor ligand binding,

signal transduction, cell adhesion, development, bacterial virulence, DNA repair and

148

RNA processing 199. The underlying similarity between these functions is molecular recognition. The LRR motif allows for protein-protein interactions which can be direct like a ligand binding component or as a regulator that affects affinity or specificity of binding to a separate ligand binding site. Binding partners to LRR have been identified

103. Probably the most significant among the binding partners of Flii identified so far is the interaction of LRR with Ras proteins which suggest a role of Flii in a signaling pathway. Therefore, proteins with LRR motif such as Flii are indicative of a multifunctional protein that includes interaction with other proteins.

4.1.3 The relationship between Flii and TGFβ

TGFβ has major roles in many cellular processes including proliferation, differentiation and apoptosis, all of which are important coordination factors in determining the outcome of the wound healing process. Classical TGFβ signals are transduced by the phosphorylation of intracellular mediators, Smad 2 and Smad 3 by the TGFβ receptors

(TβRI and TβRII) 200. The receptor activated Smads then associate with Smad 4 to form a heteromeric complex and translocate into the nucleus where they target TGFβ inducible genes 201,202. The regulation of TGFβ signaling activity is highly complex and is inbuilt with multiple controls such as the binding of cFos and cJun (AP-1) proteins to its promoter. Two distinct sites of the TGFβ promoter are responsive to regulation by the

149

AP-1 , one located at 5' upstream transcriptional start site and the other

one located between the two major start sites 203.

Instead of the classical TGFβ signaling pathway via the Smad proteins, TGFβ can also

signal independently via phosphotidylinositol 3-kinase (PI3K)/Akt, or various MAPK pathways 138,168,204. These alternate pathways act to provide additional layers of control to tightly regulate TGFβ activity. Therefore, the regulation of any of the proteins involved in the signaling pathway will affect the expression or the activity of TGFβ.

In the previous chapter, we have reported a link between TGFβ isoform expression and

wound healing in mice with varying amounts of Flii protein. Coupled with the protein

interacting capabilities of Flii protein, it is likely that Flii could be regulating TGFβ

expression and/or activity leading to a difference in the outcome of wound healing.

Therefore, we aimed to determine if Flii modulated TGFβ expression and/or signaling

by directly associating with TGFβs or its regulatory proteins.

150

4.2 Materials and Methods

4.2.1 Cells, Cell Culture

All Human Foreskin Fibroblasts (HFFs) were cultured in Dulbecco’s modified Eagle’s

Medium (DMEM) supplemented with 10% fetal calf serum (FCS) and antibiotics (100U

penicillin and 100ug/500ml streptomycin). All cell cultures were incubated at 37°C and

5% CO2. All cells were serum starved in DMEM containing antibiotics for at least 3

hours or otherwise stated prior to experimenting.

4.2.2 Antibodies

Mouse monoclonal anti-Flii antibody (sc-21716), rabbit polyclonal anti-Flii (sc-30046), anti-TGFβ1 (sc-146), anti-cFos (sc-52), anti-cJun (sc-45), anti-Akt (sc-8312) and anti-pAkt

(sc-7985) antibodies were obtained from Santa Cruz Biotechnology (CA, USA). Rabbit polyclonal anti-TGFβ2 (AB-12-NA) and goat polyclonal anti-TGFβ3 (AB-244-NA) were obtained from R&D Systems (Minneapolis, USA). Mouse monoclonal anti-Gelsolin

(610413) was obtained from BD Biosciences Pharmigen. Biotinylated horse anti-mouse

IgG from Vector (Burlingane, CA), Streptavidin Alexa Fluor 555 (S32355) and goat anti- rabbit Alexa Fluor 488 (A11008) from Invitrogen (Oregon, USA) were used in this study.

151

4.2.3 Immunocytochemistry

Cells were seeded at a density of 3 x 105 cells per well onto sterile glass coverslips and

placed in six well culture plates overnight. Cells were serum starved for 3 hours before

wounding. Cells were wounded using a P200 yellow pipette tip and incubated with growth media for a period of 30 minutes. This was followed by fixing cells with cold acetone for 10 seconds and placed back into 1x PBS containing six well plates. 1x PBS were used hereafter for every washes. 3% Normal Horse Serum (NHS) diluted in 1x

PBS were used to block cells for 30 minutes at room temperature before incubating with rabbit raised antibodies overnight at 4°C. Secondary anti-rabbit Alexa Fluor 488 were added after an initial wash for 1 hr in the dark at room temperature. This was then followed by incubating with mouse raised antibodies for 1 hr at room temperature. The respective biotinylated secondary antibodies were then added and the coverslips were incubated for 1 hr in the dark at room temperature. After this, the coverslips were incubated with streptavidin conjugated Alexa Fluor 555 for 1 hr in the dark at room temperature. DAPI was then added to the cells before a final wash and then mounted onto a microscope slide using DAKO mounting medium.

152

4.2.4 Nuclear and cytoplasmic fractionation

Preparations of nuclear and cytoplasmic fractionates were done using a nuclear extract

kit (Active Motif, Cat#40010, California, USA) according to the manufacturer’s protocol.

In a tissue culture plate of area 75cm2, growth media was aspirated from plate and washed with 5ml of ice-cold 1 x PBS containing Phosphatase Inhibitors. The solution was then discarded and a further 3ml ice-cold PBS/Phosphatase Inhibitors was added.

Cells were then removed using a cell scraper and cell suspension transferred into a yellow cap centrifuge tube. This was followed by centrifuging cells at 500rpm at 4°C and cell pellet kept on ice. Supernatant was discarded. Cells were then resuspended with 500µl 1x Hypotonic Buffer and transferred into an eppendorf tube which was then incubated for 15mins on ice. 25µl of detergent was then added and mixed thoroughly using a vortex mixer. The suspension was then centrifuged for 30secs at 14,000g at 4°C and the cytoplasmic fraction (supernatant) was transferred into a fresh eppendorf tube which could be stored at -80°C. The nuclear pellet was then resuspended in 50µl of

Complete Lysis Buffer (provided by the manufacturer) and incubated for 30mins on ice.

This was followed by mixing for 30secs and was centrifuged for 10mins at 14,000g at

4°C. The nuclear fraction was then transferred into a fresh eppendorf tube and stored at

-80°C.

153

4.2.5 Immunoprecipitation

Immunoprecipitation (IP) was used to investigate whether Flii directly interacts or associates with c-fos, c-jun, TGFβ1, TGFβ2, TGFβ, Akt, Smad 2/3 and Smad 7 in response to wounding. Nuclear and cytoplasmic fractions obtained were pre-cleared using Recombinant Protein G Agarose (Invitrogen, Cat#15920-010, Victoria, Australia) by adding 25µl rProtein G to 1ml of lysate and incubating for 10mins on a rocking platform at 4°C. rProtein G was prepared by washing stock twice in ice-cold 1x PBS prior to pre-clearing. The pre-cleared fractions were then divided into two parts of

250µl for the cytoplasmic fraction and 25µl for the nuclear fraction. Following this, 5µl of IP antibody was added to the cytoplasmic fraction and 2µl of IP antibody was added to the nuclear fraction. The samples were then incubated overnight at 4°C on a rocking platform. 50µl and 5µl of rProtein Agarose were added to the cytoplasmic and nuclear fraction respectively and incubated for a further 2 hours at 4°C on a rocking platform.

The fractions were washed 3 times with cold lysis buffer followed by centrifuging at

14,000g for 10secs and supernatant discarded. At the final wash, agarose beads were resuspended in 2x SDS Loading Buffer (25mM Tris pH 6.8, 8% Glycerol, 1%SDS and

0.02% Bromphenol blue) and were mixed thoroughly. The samples were then heated for

3mins at 95°C after which the samples were loaded onto a western gel.

154

4.2.6 Western Blotting

Protein samples were equalized by dilution and heated at 95°C prior to electrophoresis.

Protein fractions were electrophoresed on a 10% separating (3.35ml 30% Acrylamide-Bis

Solution (37.5:1, 2.6% C, BioRad Laboratories, CA, USA), 1.25ml 3M Tris pH 8.9, 5.25ml distilled water, 125ul 10% SDS, 100ul Ammonium Persulfate (APS) and 6.25µl TEMED

(N,N,N’,N’ – Tetramethylethylene-diamine, Sigma Aldrich, Sydney, Australia)) and 4% stacking (0.5ml 30% Acrylamide, 0.276ml 0.5M Tris pH 6.8, 4.104ml distilled water, 50µl

10% SDS, 40ul 10 % APS and 4µl TEMED) SDS-PAGE gels at 100V for 90 mins and then transferred onto 0.2µm pore nitrocellulose membrane (Advantec MFS Inc, CA, USA) by wet transfer (Bio-Rad Laboratories, Regents Park, NSW, Australia) using standard wet transfer - Towbin’s buffer (25mM Tris, 192mM Glycine, 20% Methanol and 0.05% SDS) at 100V for 1 hour. Membranes were stained in Ponceau Red Stain (Sigma Aldrich,

Sydney, Australia) for 10mins and then destained in distilled water and washed in PBS

Tween (0.3% Tween/1x PBS). The gel was stained in Coomassie (Sigma Aldrich, Sydney,

Australia) for 30mins and destained in destaining solution (40% Methanol, 10% Acetic

Acid and 50% distilled water) overnight.

Membranes were blocked in 5% milk blocking buffer (5% skimmed milk powder and

0.3% Tween-20 diluted in 1x PBS) for 1 hour. Primary antibodies at 1µg/ml concentration and secondary antibodies at 1µg/ml concentration were used in all

155

western blotting experiments. Primary antibodies were diluted in blocking buffer and

then added to the membrane and incubated overnight at 4°C. Stringent washes with

blocking buffer were performed every 15 minutes for an hour before appropriate

secondary antibodies conjugated with horse radish peroxidise (HRP) was added for 1

hour at room temperature. Washes were then performed before signal detection using

Super Signal West Femto Maximum Sensitivity Substrate (Pierce Biotechnology,

Rockford, USA) and signal capture using GeneSnap analysis program (Syngene,

Maryland, USA). Membranes were stripped and re-probed with β-tubulin (Sigma

Aldrich, Sydney, Australia) as a loading control.

156

4.3 Results

4.3.1 Flii associates with c-Fos and c-Jun

Using immunofluorescence staining, the localization of Flii, c-Fos and c-Jun were all visualized in unwounded and wounded fibroblasts (Figure 4.1A-F). In unwounded cells, Flii staining was localized largely to the cytoplasm and the surrounding nuclear peripheral region (Figure 4.1A, D). Both c-Fos and c-Jun were predominantly localized in the cell nucleus and the nuclear periplasmic region (Figure 4.1 B, E). However, staining for c-Fos and c-Jun was also observed in the cytoplasm. Composite images of

Flii and c-Fos staining revealed co-localization of these two proteins in both the cytoplasm and the nucleus (Figure 4.1C). This was also similarly observed for Flii and c-

Jun (Figure 4.1F). Co-localization of Flii with c-Fos/c-Jun would suggest interactions between the proteins. To confirm this observation, Flii was co-immunoprecipitated with either c-Fos or c-Jun. In addition, cellular lysates were separated into two fractions, nuclear and cytoplasmic in order to determine in which fraction Flii interacts with c-Fos and c-Jun. Interestingly, Flii co-immunoprecipitated with both c-Fos and c-Jun (Figure

4.1G). The western band intensity in the nuclear fraction was observed to be stronger than the cytoplasmic fraction. This indicated that Flii interactions with c-Fos and c-Jun were higher in the nucleus than in the cytoplasm but it could also be due to the fact that more c-Fos and c-Jun proteins are present in the cell nucleus. Gelsolin which is a

157

A Flii B c-Fos C Merge

D Flii E c-Jun F Merge

G Cytoplasmic Nuclear

c-Fos c-Jun IP: c-Fos c-Jun IP: WB: WB: Flii (145kDa) Flii (145kDa)

WB: WB: Gelsolin (88kDa) Gelsolin (88kDa)

158

Figure 4.1

Flii interacts with AP1 proteins, c-Fos and c-Jun. Dual immunofluorescence showing

protein localization of Flii and c-Fos/c-Jun. Red represents Flii staining (A, D). Green represents c-Fos and c-Jun staining (B, E). (C, F) and blue represents nuclear staining.

Merge images are composite images of Flii with c-Fos or c-Jun. Co-localizations in the cytoplasm are represented by areas of yellow color (arrow heads) stainings and nuclear co-localizations are represented by areas of pink color stainings (arrow). (G)

Immunoprecipitation showing association of Flii with c-Fos and c-Jun in cytoplasm and nucleus. n = 10. Scale bar = 50µ in (F).

159

member in the same family as Flii did not co- immunoprecipitate indicating that interaction was specific to Flii.

4.3.2 Interactions of Flii with TGFβs

The localization of Flii and all three TGFβ isoforms were determined in unwounded

and wounded human foreskin fibroblasts (HFFs). Flii was localized predominantly in

the cytoplasm as well as the nuclear peripheral area in unwounded cells. In wounded

HFFs, translocation of Flii into the nucleus was observed, as indicated by localized

staining in the cell nucleus (Figure 4.2-4.7 A, D).

TGFβ1 staining was observed to be in the cytoplasm and nucleus in unwounded HFFs

(Figure 4.2B). Staining was however noted to be stronger in the cell nucleus (Figure

4.2B). In unwounded cells, co-localization between Flii and TGFβ1 was observed in both the cytoplasm as well as in the cell nucleus (Figure 4.2C). On the other hand, wounded cells showed predominant nuclear staining for both Flii and TGFβ1 as shown in Figure

4.2 D,E. However, staining for TGFβ1 was almost entirely nuclear which indicated that wounding caused the nuclear translocation of TGFβ1 (Figure 4.2E). In the composite images of wounded cells, strong co-localization of Flii and TGFβ1 proteins were clearly

160

A Flii B TGFβ1 C Merge

Unwounded

D Flii E TGFβ1 F Merge

Wounded

G Unwounded Wounded (Cytoplasmic) (Cytoplasmic)

IP: TGFβ1 IP: TGFβ1

WB: WB: Flii (145kDa) Flii (145kDa)

WB: WB: Gelsolin (88kDa) Gelsolin (88kDa)

H Unwounded Wounded (Nuclear) (Nuclear)

IP: TGFβ1 IP: TGFβ1

WB: WB: Flii (145kDa) Flii (145kDa)

WB: WB: Gelsolin (88kDa) Gelsolin (88kDa)

161

Figure 4.2

Flii interacts with TGFβ1. Dual immunofluorescence showing protein localization of Flii and TGFβ1. (A, D) Red represents Flii staining. (B, E) Green represents TGFβ1. (C, F)

Composite image showing co-localization represented in yellow. (G)

Immunoprecipitation showing association of Flii with TGFβ1 in cytoplasm. (H)

Immunoprecipitation showing association of Flii with TGFβ1 in nucleus in unwounded and wounded cells. n = 10. Scale bar = 50µ in (F).

162

observed (Figure 4.2F). However, unlike the unwounded cells, Flii was observed to be

strongly co-localized to TGFβ1 in the nucleus with little to no co-localization in the

cytoplasm (Figure 4.2F).

To support the co-localization of Flii and TGFβ1 observations, cellular

immunoprecipitation was done. Immunoprecipitation of Flii with TGFβ1 indicated that

Flii does associate with TGFβ1, particularly in the response to wounding. Indeed, Flii

was found to immunoprecipitate with TGFβ1 in a manner consistent with previous

immunofluorescence results (Figure 4.2A-F). Flii immunoprecipitated weakly with

TGFβ1 in the cytoplasmic fractions of unwounded cells indicating the interactions of

Flii and TGFβ1 in the cytoplasm (Figure 4.2G). Wounding did not appear to increase

immunoprecipitation of Flii with TGFβ1 (Figure 4.2G). In fact, Flii co-

immunoprecipitation with TGFβ1 was not detectable. In the nuclear fractions of

unwounded cells, co-immunoprecipitation of Flii with TGFβ1 was detected in small

quantities (Figure 4.2H). This indicated that Flii also interacts with TGFβ1 in the nucleus

of unwounded cells. However, strong co-immunoprecipitation of Flii with TGFβ1 was

detected in the nuclear fraction of wounded cells (Figure 4.2H). This suggests that Flii interacts strongly with TGFβ1 in the nucleus in response to wounding.

163

Flii association with TGFβ2 was similar to TGFβ1. Localization of TGFβ2 in unwounded

HFFs was predominantly localized to the cytoplasm and the nuclear periplasmic region

(Figure 4.3B). Similarly, in response to wounding, TGFβ2 was observed to translocate

into the nucleus (Figure 4.3E). In unwounded cells, Flii co-localized with TGFβ2 mainly in the cytoplasm but co-localization was also observed in the nucleus. Upon wounding however, Flii co-localized strongly to TGFβ2 in the nucleus where both proteins were clearly present (Figure 4.3F). These observations were supported by the immunoprecipitation results (Figure 4.3G, H). Flii co-immunoprecipitated with TGFβ2 in the unwounded cytoplasmic fraction but upon wounding, less association with

TGFβ2 occurred in the cytoplasm (Figure 4.3G) due to the nuclear translocation of both proteins. Furthermore, Flii did not co-immunoprecipitate with TGFβ2 in unwounded nuclear fractions. However, as clearly shown in figure 4.3H, Flii strongly co- immunoprecipitated with TGFβ2 in the wounded nuclear fraction suggesting interaction in the cell nucleus.

A similar result was also observed with TGFβ3. Staining for TGFβ3 was observed to be present in both the cytoplasm and the cell nucleus (Figure 4.4B). Co-localization of Flii and TGFβ3 in unwounded cells was observed predominantly in the cytoplasm (Figure

4.4C). Nuclear co-localization of Flii and TGFβ3 were also observed. Wounding caused a shift in localization of TGFβ3 and Flii where staining for both proteins showed weak

164

A Flii B TGFβ2 C Merge

Unwounded

D Flii E TGFβ2 F Merge

Wounded

G Unwounded Wounded (Cytoplasmic) (Cytoplasmic)

IP: TGFβ2 IP: TGFβ2 WB: WB: Flii (145kDa) Flii (145kDa)

WB: WB: Gelsolin (88kDa) Gelsolin (88kDa)

H Unwounded Wounded (Nuclear) (Nuclear)

IP: IP: TGFβ2 TGFβ2 WB: WB: Flii (145kDa) Flii (145kDa)

WB: WB: Gelsolin (88kDa) Gelsolin (88kDa)

165

Figure 4.3

Flii associates with TGFβ2. Dual immunofluorescence showing protein localization of

Flii and TGFβ2. (A, D) Red represents Flii staining. (B, E) Green represents TGFβ2. (C,

F) Composite image showing co-localization represented in yellow. (G)

Immunoprecipitation showing association of Flii with TGFβ2 in cytoplasm. (H)

Immunoprecipitation showing association of Flii with TGFβ2 in nucleus in unwounded and wounded cells. n = 10. Scale bar = 50µm in (F).

166

A Flii B TGFβ3 C Merge

Unwounded

D Flii E TGFβ3 F Merge

Wounded

G Unwounded Wounded (Cytoplasmic) (Cytoplasmic)

IP: TGFβ3 IP: TGFβ3

WB: WB: Flii (145kDa) Flii (145kDa)

WB: WB: Gelsolin (88kDa) Gelsolin (88kDa)

H Unwounded Wounded (Nuclear) (Nuclear)

IP: TGFβ3 IP: TGFβ3 WB: WB: Flii (145kDa) Flii (145kDa)

WB: WB: Gelsolin (88kDa) Gelsolin (88kDa)

167

Figure 4.4

Flii interacts with TGFβ3. Dual immunofluorescence showing protein localization of Flii and TGFβ3. (A, D) Red represents Flii staining. (B, E) Green represents TGFβ3. (C, F)

Composite image showing co-localization represented in yellow. (G)

Immunoprecipitation showing association of Flii with TGFβ3 in cytoplasm. (H)

Immunoprecipitation showing association of Flii with TGFβ3 in nucleus in unwounded and wounded cells. n = 10. Scale bar = 50µm in (F).

168

staining in the cytoplasm whereas strong staining was observed in the cell nucleus

(Figure 4.4D, E). This indicated that both TGFβ3 and Flii translocates into the cell nucleus in response to wounding. As a result, co-localization of Flii and TGFβ3 could be clearly observed in the nucleus of wounded cells (Figure 4.4F). In agreement with the immunofluorescent results, Flii co-immunoprecipitated with TGFβ3 in the unwounded cytoplasmic fraction (Figure 4.4G). In wounded cell cytoplasmic fractions, Flii co- immunoprecipitated weakly with TGFβ3. Flii co-immunoprecipitation with TGFβ3 was also minimal in unwounded nuclear fractions but interaction between the two proteins increased significantly in response to wounding. This result reflects the observation in

Figure 4.4F where both proteins translocate into the cell nucleus in response to wounding.

4.3.3 Flii interacts with nuclear Akt in wounded cells

Akt is also an important signaling molecule involved in intracellular signaling of

TGFβs. To determine if Flii associates with Akt, co-localization and immunoprecipitation studies were performed. Flii staining was predominantly cytoplasmic in unwounded cells (Figure 4.5A). Immunofluorescent staining showed

Akt to be present in both the cell cytoplasm and nucleus in unwounded fibroblasts

(Figure 4.5B). Composite images of Flii and Akt showed co-localization of the two

169

A Flii B Akt C Merge

Unwounded

D Flii E Akt F Merge

Wounded

G Unwounded Wounded (Cytoplasmic) (Cytoplasmic)

IP: Akt IP: Akt

WB: WB: Flii (145kDa) Flii (145kDa)

WB: WB: Gelsolin (88kDa) Gelsolin (88kDa)

H Unwounded Wounded (Nuclear) (Nuclear)

IP: Akt IP: Akt

WB: WB: Flii (145kDa) Flii (145kDa)

WB: WB: Gelsolin (88kDa) Gelsolin (88kDa)

170

Figure 4.5

Flii interacts with Akt. Dual immunofluorescence showing protein localization of Flii and Akt. (A, D) Red represents Flii staining. (B, E) Green represents Akt. (C, F)

Composite image showing co-localization represented in yellow. (G)

Immunoprecipitation showing association of Flii with Akt in cytoplasm. (H)

Immunoprecipitation showing association of Flii with Akt in nucleus in unwounded and wounded cells. n = 10. Scale bar = 50µm in (F).

171

proteins in the cytoplasm of unwounded cells (Figure 4.5C). Although co-localization

was observed between Flii and Akt, the immunoprecipitation results suggest they do

not interact. No interactions between Flii and Akt were detected in the cytoplasmic and

nuclear fractions of unwounded cells (Figure 4.5G, H) indicating that Flii and Akt have

no direct interaction between them in unwounded cells.

Akt was also observed in the nucleus and cytoplasm in response to wounding (Figure

4.5E). Nuclear Akt staining appeared to be much more defined than in unwounded

cells. Composite images of Flii and Akt showed co-localization in both the nucleus and

also the cytoplasm of wounded cells. However, co-immunoprecipitation results showed

that Flii co-immunoprecipitated with Akt only in the nuclear fractions and not the cytoplasmic fractions of wounded cells indicating a nuclear association of Flii with Akt.

4.3.4 Flii associates with Smad proteins

Smad proteins are important molecules involved in TGFβ downstream signaling. Here,

Smad 2/3 and Smad 7 protein localization and any possible interactions with Flii were

investigated. Staining for Flii protein in unwounded cells was consistently localized to

the cytoplasm with some staining observed in the cell nucleus (Figure 4.6B).

Immunofluorescence staining for Smad 2/3 in unwounded cells showed localization of

172

A Flii B Smad 2/3 C Merge

Unwounded

D Flii E Smad 2/3 F Merge

Wounded

G Unwounded Wounded (Cytoplasmic) (Cytoplasmic)

IP: Smad 2/3 IP: Smad 2/3

WB: WB: Flii (145kDa) Flii (145kDa) WB: WB: Gelsolin (88kDa) Gelsolin (88kDa)

H Unwounded Wounded (Nuclear) (Nuclear)

IP: IP: Smad 2/3 Smad 2/3

WB: WB: Flii (145kDa) Flii (145kDa)

WB: WB: Gelsolin (88kDa) Gelsolin (88kDa)

173

Figure 4.6

Flii interacts with Smad 2/3. Dual immunofluorescence showing protein localization of

Flii and Smad 2/3. (A, D) Red represents Flii staining. (B, E) Green represents Smad2/3.

(C, F) Composite image showing co-localization represented in yellow. (G)

Immunoprecipitation showing association of Flii with Smad 2/3 in cytoplasm. (H)

Immunoprecipitation showing association of Flii with Smad 2/3 in nucleus in unwounded and wounded cells. n = 10. Scale bar = 50µm in (F).

174

cytoplasmic Smad 2/3 at the nuclear peripheral region. Strong Smad 2/3 staining was

also observed in the cell nucleus. Composite images revealed co-localization of Flii and

Smad 2/3 in the nuclear peripheral region of the cytoplasm in unwounded cells (Figure

4.6C). Some co-localization was also detected in the nucleus, indicating possible Flii

interactions with Smad 2/3 protein. This is consistent with the immunoprecipitation

results that showed immunoprecipitation of Flii with Smad 2/3 in both the cytoplasmic

and nuclear fractions of unwounded cells.

Wounding causes translocation of Flii protein into the cell nucleus as observed

consistently (Figure 4.6D). In wounded cells, no significant differences in Smad 2/3 localization were observed (Figure 4.6E). Smad 2/3 was still detected in both the cytoplasm and the nucleus. Composite images of Flii and Smad 2/3 showed co- localization between these two proteins in both the cytoplasm as well as in the cell nucleus (Figure 4.6F). This observation is consistent with results from the immunoprecipitation experiments. In wounded cells, Flii co-immunoprecipitated with

Smad 2/3 in both the cytoplasmic and nuclear fractions. However, stronger Flii co- immunoprecipitation with Smad 2/3 was detected in nuclear fractions of wounded cells, which indicated that Flii interacts with Smad 2/3 more strongly in response to wounding. Interestingly, in Smad pull-down results, gelsolin was found to co- immunoprecipitate with both Smad 2/3. More specifically, gelsolin was found to co-

175

immunoprecipitate with Smad 2/3 in either the cytoplasm of unwounded cells or in the

nucleus of wounded cells (Figure 4.6G, H).

Immunofluorescence for Smad 7 showed staining in both the cytoplasm and nucleus of

the cell. It was noted that strong Smad 7 staining was observed to be localized to the

nuclear peripheral space just surrounding approximately half the nuclear circumference

(Figure 4.7B). Flii staining was consistently observed to be predominantly cytoplasmic

with some staining observed in the cell nucleus (Figure 4.7A). Small amounts of Flii and

Smad 7 co-localization were observed in the region surrounding the nucleus (Figure

4.7C). This suggests that Flii may be interacting with Smad 7 in these locations. Gelsolin

interaction with Smad 7 on the other hand was evident in the nuclear fractionates of

unwounded cells (Figure 4.7H). In contrast, gelsolin did not co-immunoprecipitate with

any of the AP-1 proteins, TGFβs or Akt. However, the amount of co- immunoprecipitates were fairly low suggesting that only a limited amount of gelsolin was interacting with Smad 2/3 and Smad 7 under the described conditions.

Upon wounding, it was observed that most Smad 7 proteins were localized in and

around the cell nucleus (Figure 4.7E) which coincided with the nuclear translocation of

Flii. As a result, co-localization between Flii and Smad 7 were observed strongly in the

cell nucleus (Figure 4.7F). Cytoplasmic co-localization was also observed, which

176

A Flii B Smad 7 C Merge

Unwounded

D Flii E Smad 7 F Merge

Wounded

G Unwounded Wounded (Cytoplasmic) (Cytoplasmic)

IP: Smad 7 IP: Smad 7

WB: WB: Flii (145kDa) Flii (145kDa)

WB: WB: Gelsolin (88kDa) Gelsolin (88kDa)

H Unwounded Wounded (Nuclear) (Nuclear)

IP: Smad 7 IP: Smad 7

WB: WB: Flii (145kDa) Flii (145kDa)

WB: WB: Gelsolin (88kDa) Gelsolin (88kDa)

177

Figure 4.7

Flii interacts with Smad 7. Dual immunofluorescence showing protein localization of

Flii and Smad 7. (A, D) Red represents Flii staining. (B, E) Green represents Smad 7. (C,

F) Composite image showing co-localization represented in yellow. (G)

Immunoprecipitation showing association of Flii with Smad 7 in cytoplasm. (H)

Immunoprecipitation showing association of Flii with Smad 7 in nucleus in unwounded and wounded cells. n = 10. Scale bar = 50µm in (F).

178

suggests interaction of Flii with Smad 7 in both the cytoplasm and the nucleus.

Cytoplasmic interaction was confirmed as Flii co-immunoprecipitated with Smad 7 in both wounded and unwounded cells Figure 4.7G). Nuclear association of Flii and Smad

7 was also observed but the amount of Flii co-immunoprecipitates was almost negligible.

179

4.4 Discussion

The spatial and often transitory coordination of cytokines such as TGFβs are central to

the wound healing process and the outcome of wound repair. These mechanisms of

TGFβ signaling coordination are unclear. However, it is accepted that these processes

are controlled by complicated extracellular and intracellular signaling pathways in a

range of different cell types including fibroblasts. There are 3 isoforms of TGFβ in

mammals, all of which have a distinct role during the process of wound healing. It is clear that the regulation of specific TGFβ isoforms have to be optimal to ensure the best outcome of wound healing 116,117,120,121. The previous chapter has shown that wounds

with differential Flii gene expression levels have varying amounts of TGFβ expression.

In this chapter, the effect of wounding on Flii association with TGFβs and their

signaling molecules was investigated. The Flii protein consists of a C-terminal LRR motif allowing interactions with other proteins 102. Therefore, it was hypothesized that

Flii may be able to directly interact with TGFβs to modulate its expression or activity.

Here, we have shown that Flii not only associates with TGFβs but also proteins

involved in the regulation of expression and downstream activity of TGFβ.

Flii translocation into the cell nucleus has previously been reported with various stimuli

such as the addition of fetal calf serum or the presence of steroid hormones 109,188.

180

Immunofluorescent staining of Flii in this chapter has shown that wounding of human

foreskin fibroblasts (HFFs) in vitro also results in a similar nuclear translocation response. This indicates that Flii has functions other than just as an actin remodeling protein. It also further implicates Flii in nuclear functions. In fact, Flii has been identified as a nuclear co-activator involved in nuclear receptor signaling for thyroid

and estrogen hormone receptors 81. In the nucleus, Flii was shown to interact directly

with nuclear receptors and their co-activators, CARM1 and CBP (p300) through GRIP1

(Figure 1) 105,106. Other members of the gelsolin family, including gelsolin 81 and

supervillin 176,177 have also been identified as nuclear receptor co-activators. In response to wounding, Flii may also have co-activating roles in the nucleus.

This study has shown that Flii associates with AP-1 proteins, c-Fos and c-Jun. Since AP-

1 proteins bind to the promoter of TGFβs to regulate its expression, interactions with

Flii could be one way that Flii may affect TGFβ gene expression. As the β-ZIP region of c-Jun interacts directly with CBP which leads to transactivation 205, it is therefore

possible that Flii may interact with c-Fos and c-Jun and/or other co-activator complexes including CBP on TGFβ promoter to modulate TGFβ expression. It is important to note that TGFβs can also regulate the expression of AP-1 proteins 206. Considering that the

expression of both TGFβ and AP-1 proteins are involved in feedback loops that

181

Brg1 ATPase CARM1 Arp4 Methyl Actin Flii LRR Flii Gelsolin GRIP1 CBP (p300) Domain Domain (p160)

Acetyl NR NR Histone DNA H3

NR Response Element

Figure 1

Theoretical depiction of Flii interactions on a NR responsive gene promoter. Enzymatic activities are indicated by the arrows. NR denotes nuclear receptor. The complex depicted is theoretical since interactions are not necessary simultaneous.

182

indirectly regulate one another’s expression, Flii association with TGFβs was also determined. Direct association of Flii with TGFβs would suggest that Flii could either positively or negatively regulate the activities of TGFβ. Indeed, Flii association with all three TGFβ isoforms were detected throughout the cell. The detection of nuclear TGFβs in this study was unexpected since TGFβs are molecules which bind to receptors that transduce signals to the cell nucleus. However, nuclear TGFβs has been reported to be present in the cell nucleus 207. Therefore, the finding of nuclear TGFβ is specific and not likely an experimental artifact. Interestingly, Flii was found to interact strongly with all isoforms of TGFβ in the nucleus of wounded HFFs. The role of Flii and TGFβ nuclear interactions remains unknown but it suggests a possible regulation mechanism by Flii.

Given this, Flii interaction with TGFβ may function as another level of control during the signaling process which also suggests that the activity of TGFβ may be directly regulated by Flii.

To further investigate the effects of Flii on TGFβ functions, Smad proteins were investigated as these are the downstream signaling molecules that transduce TGFβ signals. Smad 2/3 is activated following phosphorylation by TGFβ bound receptors and consequently complexes with Smad 4 to translocate into the nucleus 134. This process is regulated by inhibitory Smad 7 which prevents the activation of Smad 2/3, thereby

183

arresting the signaling pathway 208. Flii association with Smad 2/3 was detected in both

wounded and unwounded cells. In fact, most association between Flii and Smad 2/3

occurred in the nucleus of wounded cells. Since Smad 2/3 function as co-activators in the cell nucleus, it is logical that if Flii has co-activating roles in wound healing that it interacts with nuclear Smad 2/3 to regulate TGFβ induced signaling.

It has been shown that Smad 2 and Smad 3 interact with proteins containing LRR 209.

One of the proteins, Erbin (ErbB2/Her2-interacting protein), a regulator of tyrosine

kinase function has been shown to directly interact with Smad 2 and Smad 3 proteins

210,211. Dai et al (2007) have reported that Erbin inhibits TGFβ signaling by physically

sequestering Smad 2/3 from their association with Smad 4. As a result, Erbin negatively

regulates the downstream TGFβ signaling pathway, thereby affecting transcriptional

responses. Considering that Flii contains 16 tandem LRR motifs and also interacts with

Smad 2/3, it is tempting to speculate that Flii may function to affect TGFβ signaling in a

similar fashion as Erbin. In other words, the nuclear interactions between Flii and Smad

2/3 in wounded HFFs may also sequester Smad 2/3 from Smad 4 therefore preventing

Smad 4 association with other transcriptional factors and inhibiting target gene

expression. This may also explain why most association was detected in the nucleus of

wounded cells.

184

On the other hand, Flii association with Smad 7 was also detected in low quantities in both the cell cytoplasm and nucleus. Nuclear interactions between Flii and Smad 7 were almost negligible which is expected because Smad 7 mostly functions as a cytoplasmic inhibitory protein. This does not imply that Smad 7 is an exclusively cytoplasmic protein as studies have shown localization of Smad 7 in the cell nucleus as well as the presence of a nuclear localization signal domain 208,212. The lack of interactions between

Flii and Smad 7 compared to Smad 2/3 would suggest that the regulation of TGFβ by

Flii is most likely to be in the nucleus where Flii co-activate gene transcription with

Smad 2/3 for cellular growth rather than the regulation by sequestering or activating inhibitory Smad 7.

Flii association with Akt was also investigated in this study due to the convergence of the map kinase and TGFβ signaling pathways 213,214. In addition, Akt is also able to regulate TGFβ signaling by physically binding and activating Smad 2/3 through the map kinase signaling pathway, which then enables Smad 2/3 to form complexes and target response genes 215,216. In this study, Flii was found to associate with Akt only in the nucleus of wounded cells. No interaction between Flii and Akt was detected in the cytoplasm of wounded cells or in the cytoplasm and nucleus of unwounded cells. The specificity of Flii and Akt interaction in the cell nucleus would imply a unique role for

Flii in Akt function such as cellular survival or the regulation of the cell cycle 217-219. This

185

result further strengthens Flii function in cellular signaling. Taken together, the results

indicated that Flii may have roles as a modulator in the TGFβ signaling pathway.

In summary, the novel findings of Flii interaction with multiple proteins involved in

TGFβ expression and activity indicated a mechanistic role for Flii in TGFβ regulation. In

the present study, Flii interacting proteins including c-Fos and c-Jun, TGFβ1, TGFβ2,

TGFβ3, Smad 2/3, Smad 7 and Akt were identified. These proteins are all regulatory

“check-points” which determine the effects of TGFβ. It is also interesting to see that these interactions occur most strongly in the nucleus of wounded cells. These findings

provided a clue in the regulation of TGFβ by Flii in the cell nucleus. Therefore, it is

likely that Flii may either act as a nuclear repressor or as a co-activator to influence the

activity of TGFβ on target gene expression. The underlying mechanism of TGFβ

regulation by Flii remains unclear and will be investigated further in the next chapter.

Nevertheless, Flii may regulate wound repair directly through its interaction with TGFβ

or the TGFβ regulatory proteins which include the AP-1 proteins, Akt and Smads.

186

CHAPTER FIVE

Regulation of TGFβ by Flii association with the MAPK

signaling pathway

5.1 Introduction

5.1.1 Interaction of Flii with Ras GTPases

The LRR motif in Flii is similar to a subgroup of LRR containing proteins that interacts

with Ras ligands 197. Among these proteins are human and mouse Rsp-1 involved in

suppressing v-Ras transformation of cells and the membrane associated yeast

(Saccharomyces cerevisae) adenylate cyclase in which LRR motifs are a prerequisite for

interaction with Ras proteins. Interestingly, the Flii LRR motif has approximately 35%

identity and 53% similarity to Rsp-1 220. Taken together, there is a strong possibility that

Flii may resemble Rsp-1 and have similar roles in Ras signal transduction 199. An

interaction between Flii and Ras was first identified in migrating Swiss 3T3 fibroblast 109.

In addition to Ras, other small GTPases that interact with Flii were also identified,

which includes RhoA and Cdc42. Coupled with the fact that phosphoinositide 3-kinase

can regulate gelsolin and other actin binding proteins, Flii may be involved in signaling

pathways. This is likely given that Ras is a part of the MAPK network and therefore

interaction with Ras could affect MAPK signaling networks.

187

The finding that Flii can act as a signaling molecule has only been discovered recently

and Flii has been implicated in various signaling processes such as co-activating nuclear hormone receptors, inflammation and cell survival, which affects wound healing 82,107,108.

Current research activity is centered on Flii capabilities of interaction with other

proteins through its LRR domain to affect target gene responses 105,107,198. Similarly,

previous chapters have shown direct interactions between Flii and TGFβ as well as

proteins involved in TGFβ signaling pathways, therefore implicating Flii in the

regulation of TGFβ expression. It is yet to be established how Flii is integrated into the

signaling networks of TGFβ. There are multiple signaling pathways that can affect

TGFβ signaling in mammalian cells, including the MAPK signaling pathway 120,221.

Previous findings on Flii interaction with Ras proteins suggest that Flii is linked to the

MAPK signaling pathway 109. This chapter investigates the underlying mechanism of

Flii regulation of TGFβ in more depth by using MAPK inhibitors in cells derived from

WT, Flii+/- and FliiTg/+ mice.

188

5.2 Materials and Methods

5.2.1 siRNA knockdown

HFFs were seeded into 6 well tissue culture plates and cultured overnight to achieve

30% to 50% confluence at time of transfection. Sequence of Flii siRNA are as follows;

forward: 5’-GCU GGA ACA CUU GUC UGU GTT-3’, reverse: 5’-CAC AGA CAA GUG

UUC CAG CTT-3’ 105. siRNA were transfected into the cells using Lipofectamine 2000

(Invitrogen, Carlsbad, USA). Both siRNA and Lipofectamine 2000 were diluted in Opti-

MEM I Reduced Serum Medium (Invitrogen, Carlsbad, USA). 250µl of siRNA

(optimized to 100nM per well) was mixed with 4µg of Lipofectamine 2000 diluted in

250µl Opti-MEM and were allowed to complex at room temperature for 20 minutes.

500µl of siRNA:Lipofectamine 2000 complex was then added to each well, mixed and

cells incubated for 6 hours before replacing transfection media with DMEM containing

10% FCS only. Cells were incubated for 48 hours prior to gene knockdown assessment.

5.2.2 RNA extraction

RNA was extracted from Human Foreskin Fibroblasts. RNA extraction from cell

cultures only required scraping cells from culture flasks after adding Trizol Reagent

(Invitrogen, Victoria, Australia). Samples were transferred into fresh eppendorf tubes

and centrifuged at 12,000g at 4°C for 10mins to remove cell debris. The samples were

189

incubated for 5mins at room temperature before adding 200µl of chloroform to each

tube and were mixed thoroughly by hand for 15secs. The samples were kept at room

temperature for 3mins and centrifuged at 12,000g for 15mins at 4°C. The aqueous phase

containing RNA was transferred into a fresh tube and 500µl of isopropanol added to

precipitate the RNA. The samples were then incubated at room temperature for 10mins

before centrifuging at 12,000g for 10mins at 4°C. The supernatant was discarded and the

residual pellet washed with 1ml of 75% ethanol. Finally, samples were centrifuged at

7500g for 5mins at 4°C and supernatant discarded. The pellet was dried and re- dissolved in 50µl DEPC water.

5.2.3 DNase Treatment and RNA Quantitation

RNA samples obtained were subjected to DNA-free DNase Treatment and Removal Kit

(Ambion, TX, USA) as instructed by the manufacturer to remove any contaminating

genomic DNA. Firstly, RNA samples were treated with 0.1 volume of 10x DNase Buffer

and 1µl of rDNase I and incubated at 37°C for 30mins. Following this, 0.1 volume of

DNAse Inactivating Reagent was added to the samples and incubated at 2mins at room

temperature with occasional mixing. The samples were then centrifuged at 10,000g for

90secs and supernatant transferred to fresh tubes. RNA was quantitated by diluting 1 in

20 with RNase free water and 100µl duplicates were quantified using a Pharmacia

Biotech GeneQuant RNA/DNA Calculator using RNase-free water as a blank.

190

Absorbance at 260nm and 280nm were measured that quantify RNA absorbance as

µg/µl concentration. Purity of RNA was confirmed by the A260/A280 ratio and a value between 1.7 to 2.0 indicates good RNA quality.

5.2.4 Complementary Deoxyribonucleic Acid (cDNA) Synthesis cDNA was synthesized from RNA using reverse transcription. Each reaction contains

1µg of RNA with 4µl 2.5µM dNTPs (dATP, dCTP, dGTP and dTTP, 100mM each,

Promega, WI, USA) and 2µl Oligo(dt)12-18 Primer (25µg at 0.5µg/µl, Invitrogen, Victoria,

Australia). This was heated at 85°C for 3mins and placed in ice immediately. 2µl 10x

Stratascript Buffer (Stratagene, Epson, UK), 1µl RNasin (Promega, WI, USA) and 1µl

Stratascript Reverse Transcriptase (Stratagene, Epson, UK) were added to the mixture and heated at 42°C for 60mins followed by 92°C for 10mins before cooling it down on ice. A control sample was prepared with the reagents above with the exclusion of reverse transcriptase for use in the Real-Time quantitative-Polymerase Chain Reaction

(RTq-PCR) as a negative control.

5.2.5 Real-Time quantitative-Polymerase Chain Reaction (RTq-PCR)

Each PCR reaction tube containing cDNA was set up to a final concentration of 1x SYBR

Green, 1x Amplitaq PCR buffer, 3mM MgCl2, 5mM dNTPs, 0.9µM primers (forward

191

and reverse), 1.25 Units of Amplitaq Gold DNA polymerase in 25µl of H2O. Primer sequences used in this chapter is shown in Table 3.2. The cycle conditions are as follows; an initial denaturation at 95°C for 15 minutes, 35 cycles of 95°C for 25 seconds, 60°C for

30 seconds, 72°C for 30 seconds and at the final cycle, an additional 5 minutes at 72°C before a melt from 72°C to 99°C at 30 seconds at each degree step. Refer to Table 1 for primer sequence.

5.2.6 Primary Fibroblast Extraction and Culture

Refer to Section 3.3.1 – 3.3.4 for information on the generation of Flii+/- and FliiTg/+ mice

and ethics for surgery to obtain skin biopsies. Full thickness unwounded skin was

removed from the dorsum of WT, Flii+/- and FliiTg/+ mice and placed separately in ice-

cold Ham’s Nutrient Mixture F12 cell media (SAFC Biosciences, Kansas, USA)

supplemented with 5% Penicillin Streptomycin (Sigma Aldrich, Sydney, Australia), 5%

Fungizone (Sigma Aldrich, Sydney, Australia). 2mm punch biopsies (Acupunch,

ACUDERM) were then taken and fixed dermal side down onto 6 well tissue culture

plates. 4 punch biopsies were seeded per well. The explants were allowed to air-dry for

at least 30 minutes until attached firmly to the culture wells. The explants were then

cultured in DMEM cell culture media containing 20% FCS, 5% Penicillin Streptomycin,

5% Fungizone at 37°C and 5% CO2. Culture media was replaced 24 hours later and then

every two days. Fibroblasts outgrowth was observed after two days. The explants were

192

Forward or Primer Reverse Sequence 5' - 3' forward CCT CCT ACA GCT AGC AGG TTA TCA AC Flii reverse GCA TGT GCT GGA TAT ATA CCT GGC AG forward CAG ACA GCC CCT GCC AGC ACC C Gelsolin reverse GAG TTC AGT GCA CCA GCC TTA GGC forward GGT TGG ATG GCA AGC ATG TG Cyclophillin A reverse TGC TGG TCT TGC CAT TCC TG forward CTA CGA GGC GTC ATC CTC CCG c-Fos reverse AGC TCC CTC CGG TTG CGG CAT forward GAA ACG ACC TTC TAT GAC GAT GCC CTC AA c-Jun reverse GAA CCC CTC CTG CTC ATC TGT CAC GTT CTT forward GTT GGA CGA GCT GGA GAA GG Smad 3 reverse TGC TGT GGT TCA TCT GGT GG forward ACG GCC ATC TTC AGC ACC AC Smad 4 reverse AGA ATG CAC AAT CGC CGG AG forward GCT CAC GCA CTC GGT GCT CA Smad 7 reverse CCA GGC TCC AGA AGA AGT TG

Table 1

Primer sequence used in RT-qPCR.

193

removed from the wells at day 7 and the primary fibroblasts were allowed to close the

gap where the explants originally occupy. At 90% cell confluence, the fibroblasts were

washed twice using sterile 1X PBS and trypsinized with 1X Trypsin diluted in DMEM.

The fibroblasts were then seeded into 75cm2 tissue culture flasks (3 wells to 1 flask) and

cultured to confluence in 20% FCS DMEM cell culture media stated above. This would

be the primary cell stocks and were used to propagate fibroblasts population until

passage 10.

5.2.7 Immunocytochemistry

Primary WT, Flii+/- and FliiTg/+ fibroblasts were seeded at a density of 3 x 105 cells per

well onto sterile glass coverslips and placed in six well culture plates overnight. Cells

were serum starved for 3 hours followed by fixing cells with cold acetone for 10 seconds

and placed back into 1x PBS containing six well plates. Washes were performed after

every treatments using 1x PBS. 3% Normal Horse Serum (NHS) diluted in 1x PBS were

used to block cells for 30 minutes at room temperature before incubating with anti α-

SMA antibodies (1:200 dilution) overnight at 4°C. Secondary biotinylated anti-mouse was then added at a dilution of 1:1000 for 1 hr in the dark at room temperature. After this, streptavidin conjugated Alexa Fluor 555 at 1:200 dilution was added for 1 hr in the dark at room temperature. DAPI was added to the cells at 1:1000 dilution before a final wash and then mounted onto a microscope slide using DAKO mounting medium.

194

5.2.8 Collagen Secretion Assay

Collagen secretion in WT, Flii+/- and FliiTg/+ explants and primary fibroblasts were

quantitated using Sircol Collagen Assay Kit according to the manufacturer’s protocols.

Spent media collected from confluent cell samples were centrifuged at 15,000 x g for 30

minutes and supernatant collected to remove any cell debris. A standard curve

containing 0, 50, 125 and 250µg/ml of collagen and test samples were prepared. 1ml of

Sircol dye reagent was added to all samples and mixed thoroughly before incubating at

room temperature for 30 minutes on an orbital rotator. This was followed by

centrifuging at 12,000 x g for 10 minutes at room temperature and supernatant

discarded after centrifugation. 1ml of Alkali reagent was added and mixed thoroughly

to completely dissolve the pellet. Sample volumes of 200µl were transferred to 96

microplate wells and absorbance measured at 540nm.

5.2.9 Proliferation Assay

WST-1 proliferation reagent (Cayman Chemical, WI, USA) was used to assess the rate of

proliferation of primary cell cultures. This is a tetrazolium salt based reagent that converts to soluble formazan by dividing cells therefore directly correlates to cellular

metabolic activity and their proliferation rate. Primary WT, Flii+/- and FliiTg/+ fibroblasts

were seeded into 96-well microtitre plates at a density of 105 cells per well in 100µl of

195

10%FCS DMEM with antibiotics. Cells were then incubated at 37°C CO2 incubator

overnight and serum starved for 3 hours the following day. Cells were then treated with

1000pg/ml of TGFβ1 or MAPK inhibitors (50µM), UO126, PD98059, SP600125 or PI3K

inhibitor, LY694002 and incubated for 48 hours and assayed by adding 10µl of WST-1 reagent and mixed thoroughly for 1 min on an orbital shaker. The cells were then incubated at 37°C for 1 hour as per manufacturer’s protocol before measuring dual absorbance at 450nm and 600nm on a Tecan microplate reader. Results are normalized to the untreated control sample and expressed as a percentage.

5.2.10 Fibroblast Outgrowth Assay

2mm punch biopsies from WT, Flii+/- and FliiTg/+mice skin were collected as described in

Primary Fibroblast Preparation and Culture. 2 biopsies were seeded into each well of a

12 well culture plate and cultured in DMEM supplemented with 20% FCS, 5% Penicillin

Streptomycin, 5% Fungizone at 37°C and 5% CO2 overnight. After the initial overnight

incubation, treatments were added to the explants. 1000pg/ml of TGFβ and/or 50µM of

MAPK inhibitors were added to the explants. Cellular growth media including

treatments were replaced every second day. Random photographs were taken, 3 images

per explant, 9 explants per treatment group per mice group on a daily basis until day 6

where the farthest outgrowth is just within the field of view. Proximal outgrowth

distance was measured and recorded using AnalySIS software.

196

5.2.11 Statistical Analysis

All statistical differences were determined using the Student’s t-test or ANOVA. P value of less than 0.05 was considered significant.

197

5.3 Results

5.3.1 Flii gene knockdown affects expression of AP-1 and Smad proteins

To substantiate that Flii can regulate TGFβs, Flii knockdown was performed and expression of TGFβ related genes such as cFos, cJun, Smad 3, Smad 4 and Smad 7 were determined. Flii gene expression was knocked down using Flii specific siRNA in normal human foreskin fibroblasts (HFFs). Expression was then quantitated using real- time quantitative polymerase chain reaction. Flii gene expression levels after knockdown were approximately 22% which was significantly lower than normal levels

(refer to Chapter 3, Figure 3.8). Gelsolin gene expression was included as a specificity control to show the specificity of Flii siRNA knockdown and that gelsolin gene levels did not increase to compensate for the loss of Flii expression. When Flii gene expression was reduced, cFos and cJun gene expressions were found to be significantly lower, at approximately 55% of control (Figure 5.1A). Smad proteins which are involved in TGFβ downstream signaling are also affected in response to the reduction in Flii expression levels. Smad 3 and 4 which are responsible for the signal transduction of TGFβ had significantly lower expression (Figure 5.1B). However, the opposite was found for Smad

7 where its gene expression was significantly elevated. Over-expression of Flii in HFFs was attempted but transfection of a Flii vector for over-expression into HFFs proved to be too difficult. Transfection was either not successful or it produced significant cell

198

A

120

100

80 * * 60 40 *

20 Fold Change (% Control) to Change Fold

0 Flii Gelsolin c-Fos c-Jun

B 180 * 160

140 120 100 80 * * 60 40 *

Fold Change (% Control) to Change Fold 20 0

Flii Gelsolin Smad 3 Smad 4 Smad 7

199

Figure 5.1

Effects of Flii siRNA knockdown on cFos, cJun, Smad 3, Smad 4 and Smad 7 gene

expression. Flii gene expression was knocked down using siRNA in normal human

foreskin fibroblasts. Gene expression of AP-1 proteins and Smads were quantitated using RTq-PCR. Results are expressed as a percentage of control which is represented by the dotted line at 100%. Gelsolin, a member of the family is included as a control gene to show target specificity of Flii siRNA. Cyclophillin was used as the reference gene. n = 6, *p < 0.05.

200

death. This limited the analyses of in vitro Flii manipulation to just the siRNA

knockdown approach.

5.3.2 TGFβ1 increases collagen secretion in WT, Flii+/- and FliiTg/+ primary fibroblasts

Collagen is a component of the extra cellular matrix and is crucial in wound healing and

is produced by myofibroblasts which was also investigated in this study. Although

there were very little differences in α-SMA expression between WT, Flii+/- and FliiTg/+

fibroblasts, the amount of collagen secreted by these fibroblasts was significantly

different. In unwounded cells, it was found that WT fibroblasts secrete the least amount

of collagen followed by Flii+/- and FliiTg/+ fibroblasts, which secrete the most collagen

(Figure 5.2A). Wounding caused increased production of collagen. WT and FliiTg/+

collagen secretion increased significantly but interestingly, there was no difference in

collagen secretion by Flii+/- fibroblasts in response to wounding. The collagen levels secreted by Flii+/- fibroblasts were significantly lower than both unwounded and wounded FliiTg/+ fibroblasts but higher compared to WT fibroblasts. The effects of

TGFβ1 addition on collagen production was also studied since TGFβ1 induces

fibroblast differentiation into myofibroblasts which are the primary producers of

collagen. Adding TGFβ1 considerably increased collagen secretion in all WT, Flii+/- and

201

A Unwounded Untreated Wounded * 1500 *

* 1000 *

500 Collagen (ug/ml)

0 Flii+/- WT FliiTg/+

B Unwounded + TGF-β1 Wounded * * 3000

2000

1000

Collagen (ug/ml) 0 Flii+/- WT FliiTg/+

202

Figure 5.2

Comparison of collagen content secreted by unwounded and wounded fibroblasts.

Spent growth media from confluent, scratched wounded and TGFβ1 treated Flii+/-, WT and FliiTg/+ fibroblasts were collected and assayed for soluble collagen content. (A)

Comparison of collagen secretion in unwounded and wounded primary fibroblasts. (B)

Comparison of collagen content in TGFβ1 treated unwounded and wounded primary fibroblasts. Results represent mean ± S.E.M. n = 6, p < 0.05.

203

FliiTg/+ fibroblasts (Figure 5.2B). WT fibroblasts secreted the least amount of collagen

followed by Flii+/- fibroblasts and FliiTg/+ fibroblasts in response to TGFβ1 addition.

Wounding and the addition of TGFβ1 further increased the production of collagen in all

WT, Flii+/- and FliiTg/+ fibroblasts. FliiTg/+ fibroblasts had the least increase in collagen levels which may be the limit due to the already high collagen levels. However, it was found that TGFβ1 increased Flii+/- fibroblast collagen production to be comparable to

FliiTg/+ fibroblasts.

5.3.3 TGFβ1 decreases FliiTg/+ fibroblast proliferation

The effects of TGFβ1 on WT, Flii+/- and FliiTg/+ primary fibroblasts proliferation were

investigated to determine if exogenous TGFβ1 could affect cells with differential Flii

gene expression. Primary fibroblasts derived from WT, Flii+/- and FliiTg/+ mice were treated with 500pg/ml, 1000pg/ml and 2000pg/ml of TGFβ1 ligand. It was noted that

Flii+/- primary fibroblasts had the highest proliferation followed by WT and FliiTg/+

primary fibroblasts under normal conditions (Figure 5.3A). It was observed that

proliferation in WT and Flii+/- primary fibroblasts was significantly increased in response to TGFβ1 (Figure 5.3A). However, the opposite was observed in FliiTg/+

primary fibroblasts. Proliferation in FliiTg/+ primary fibroblasts was decreased in

response to TGFβ1 addition. The addition of 1000pg/ml (Figure 5.3B) and 2000pg/ml

(Figure 5.3C) of TGFβ1 ligand had a similar result. WT and Flii+/- primary fibroblasts

204

A 500pg/ml 250 * 200

150 * Untreated 100 +TGFβ1

Proliferation * 50 (% to WT (% to WT Untreated) 0 Flii+/- WT FliiTg/+

B 1000pg/ml 250 * 200

150 * Untreated 100 +TGFβ1 Proliferation * 50 (% to WT (% to WT Untreated) 0 Flii+/- WT FliiTg/+

2000pg/ml C 250 * 200

150 * Untreated 100 +TGFβ1 Proliferation * 50 (% to WT (% to WT Untreated) 0 Flii+/- WT FliiTg/+

205

Figure 5.3

TGFβ1 increases proliferation in Flii+/- fibroblasts but decreases proliferation in FliiTg/+

fibroblasts. Cells were serum-starved prior to adding 500pg, 1000pg and 2000pg of

TGFβ1 ligand. Proliferation status was assayed by using WST-1 reagent and absorbance

measured at 450nm and 600nm at 48 hours post-treatment. (A) Primary fibroblasts treated with 500pg/ml of TGFβ1. (B) Primary fibroblasts treated with 1000pg/ml of

TGFβ1. (C) Primary fibroblasts treated with 2000pg/ml of TGFβ1. Results represent mean ± S.E.M. n = 6, *p < 0.05 (untreated vs +TGFβ1 within each group).

206

consistently showed increased proliferation and FliiTg/+ primary fibroblasts had decreased proliferation.

5.3.4 Proliferation of primary fibroblasts treated with MAPK inhibitors

Four pharmaceutical inhibitors (UO126, PD98059, SP600125 and LY294002) were used to investigate the effects of signaling pathways on cellular proliferation in primary fibroblasts derived from WT, Flii+/- and FliiTg/+ mice expressing different levels of Flii protein. Firstly, the concentrations of inhibitors used in this study were optimised using

HFFs and inhibitor concentrations of 0µM (untreated), 10µM, 30µM and 50µM (Figure

5.4). MAPK inhibitors UO126 and PD98059 inhibit the MEK1/2 and Raf signaling pathway respectively, which resulted in the decrease of pERK. This indicated that the inhibitors were effective at starting from 30µM up to 50µM . Treatment with JNK inhibitor and PI3K inhibitor decreased Jun and pAkt expression with increasing inhibitor doses indicating the inhibitors are functioning. However, even at 50µM concentrations, JNK and PI3K inhibitors did not work as well as UO126 and PD98059 inhibitors. Increased concentrations were used but resulted in significant cell death. As a result, the highest and most effective inhibitor concentration of 50µM was used in this study.

207

Untreated 10µM 30µM 50µM

MEK1/2 Inhibitor IB: pERK (UO126)

Raf Inhibitor IB: pERK (PD98059)

JNK Inhibitor IB: Jun (SP600125)

PI3K Inhibitor IB: pAkt (LY294002)

Figure 5.4

Optimization of MAPK and PI3K inhibitors in HFFs. Cells were seeded at 50% confluency overnight and serum starved the following day. HFFs were treated with

10µM, 30µM or 50µM of MAPK (UO126, PD98059, SP600125) or PI3K (LY294002) inhibitors for 48 hours before protein samples were obtained. Untreated samples were included in the western blot to show original expression prior to the treatments.

208

In untreated cells, proliferation was the highest in Flii+/- primary fibroblasts followed by

WT and FliiTg/+ primary fibroblasts (Figure 5.5). In MEK 1/2 inhibitor UO126 treated

cells, Flii+/- and WT primary fibroblasts showed increased proliferation whereas FliiTg/+

fibroblasts had decreased proliferation (Figure 5.5A). In cells treated with Raf inhibitor

PD98059, no significant increase in proliferation was observed in Flii+/- fibroblasts

(Figure 5.5B). WT fibroblasts showed significantly increased proliferation and FliiTg/+

fibroblasts had decreased proliferation. In cells treated with JNK/SAPK inhibitor

SP600125, a significant decrease in proliferation was observed in Flii+/- fibroblasts

(Figure 5.5C). Proliferation in WT fibroblasts was still increased. FliiTg/+ fibroblasts

however, did not show any significant differences in response to SP600125 treatment. In

cells treated with PI3K inhibitor LY294002, increased proliferation was observed in WT

and Flii+/- fibroblasts but FliiTg/+ fibroblasts had decreased proliferation.

5.3.5 TGFβ1 decreases migration in Flii+/- fibroblasts

Using an outgrowth model where the distance from the explant to the farthest fibroblast

was measured. FliiTg/+ fibroblasts noticeably had slower migration compared to WT and

Flii+/- fibroblasts at day 5 and day 6 (Figure 5.6A). At the end of day 6, Flii+/- fibroblasts

were observed to migrate the farthest. There was no difference in migration between

WT and FliiTg/+ fibroblasts at day 6. The effects of TGFβ1 on migration of primary

fibroblasts were also investigated. It was found that migration of Flii+/- fibroblasts was

209

Untreated Untreated A UO 126 B PD 98059 UO 126 PD 98059 250 200 * 200 150 * 150 * 100 100 * * 50 Proliferation Proliferation 50 (% to WT Untreated) WT to (% (% to WT Untreated) WT to (% 0 0 Flii+/- WT FliiTg/+ Flii+/- WT FliiTg/+

Untreated Untreated C SP 600125 D LY 294002 SP 600125 LY 294002 150 200 * * 150 100 * * 100 50 * 50 Proliferation Proliferation (% to WT Untreated) WT to (% Untreated) WT to (% 0 0 Flii+/- WT FliiTg/+ Flii+/- WT FliiTg/+

210

Figure 5.5

Differential effects of MAPK inhibitor treatments in WT, Flii+/- and FliiTg/+ primary fibroblasts. Primary fibroblasts were seeded at 50% confluency overnight and serum- starved the following day. All inhibitors PD98059, UO126, LY294002 and SP600125 were added for 48 hours after serum-starvation and proliferation status quantitated and expressed as a percentage to untreated WT controls. (A) Primary fibroblasts treated with MEK 1/2 inhibitor UO 126. (B) Primary fibroblasts treated with Raf inhibitor PD

98059. (C) Primary fibroblasts treated with JNK/SAPK inhibitor SP 600125. (D) Primary fibroblasts treated with PI3K inhibitor LY 294002. Results represent mean ± S.E.M. n = 6,

*p < 0.05 (untreated vs +TGFβ1 within each group).

211

A Outgrowth B Flii+/- 300 * 300 * Flii+/- Untreated 250 250 WT +TGF-b1 200 * 200 * FliiTg/+ 150 150 100 100 50 50 0 0 Migration Distance Distance Migration (um) Migration Distance Distance Migration (um) 1 2 3 4 5 6 1 2 3 4 5 6

Days Days *c

C WT D FliiTg/+

300 250 * Untreated * Untreated 250 200 +TGF-b1 +TGF-b1 * 200 150 150 * 100 100 50 50 Migration Distance Distance Migration (um) Distance Migration (um) 0 0 1 2 3 4 5 6 1 2 3 4 5 6 Days Days

212

Figure 5.6

TGFβ1 decreases Flii+/- outgrowth motility. Full thickness skin explants biopsied from

WT, Flii heterozygous and over-expressing mice were seeded into culture plates.

Growth media containing 1ng/ml of TGFβ1 was then added and cellular outgrowth from explants measured daily for 6 days. (A) Comparison of outgrowth distance of

Flii+/-, WT and FliiTg/+ explants. (B) TGFβ1 inhibited outgrowth migration in Flii+/- explants. (C) TGFβ1 increased outgrowth migration in WT explants. (D) TGFβ1 inhibited outgrowth migration in FliiTg/+ explants. Results represent mean ± S.E.M. n = 6,

*p < 0.05 vs WT (A) or Untreated (B-D).

213

significantly decreased at day 5 onwards in response to TGFβ1 treatment (Figure 5.6B).

Conversely, TGFβ1 was observed to slightly increase migration in WT at day 6 (Figure

5.6C). TGFβ1 significantly increased migration in FliiTg/+ fibroblasts from day 4 to day 6

(Figure 5.6D).

5.3.6 MAPK inhibitors decrease fibroblasts outgrowth

Interactions between Flii and the MAPK pathway to regulate TGFβ has not been

studied before but will be investigated in this study using specific MAPK inhibitors and

the addition of TGFβ1. Firstly, the effects of inhibitor addition on WT fibroblasts

migration were determined. All inhibitors generally decreased outgrowth migration

(Figure 5.7A). MEK 1/2 inhibitor UO126 and Raf inhibitor PD98059-treated explants

showed significantly lesser outgrowth compared to the untreated explants, especially

from day 4 to day 6. There was no considerable difference in outgrowths in explants treated with inhibitors UO126 and PD98059. However, in explants treated with

JNK/SAPK inhibitor SP600125, fibroblast migration was not observed until day 6. This was also observed similarly in explants treated with PI3K inhibitor LY294002. The lack of outgrowth in these explants treated with SP600125 and LY294002 suggests that the

JNK/SAPK and PI3K signaling pathway is important in cellular migration.

214

A WT Explants

300 Untreated 250 SP600125 LY694002 *w-ac 200 UO126

PD98059 150 *o-v

100 *g-n

*a-f Migration Distance Distance Migration (um) 50

0 1 2 3 4 5 6

Days

B WT Explants

300 Untreated 250 SP + b1 LY + b1 *t-x 200 UO + b1 150 PD + b1 *m-s *g-l 100 *a-f

Migration Distance Distance Migration (um) 50

0 1 2 3 4 5 6

Days

215

Figure 5.7

WT explants treated with MAPK inhibitors and TGFβ1 ligand. Full thickness skin

biopsies were taken from WT mice. Explants were cultured for a period of 6 days and

cellular outgrowth measured daily. (A) WT explants treated with MAPK inhibitors

only. Explants treated with either inhibitor UO126 or PD98059 had reduced cell

migration while inhibitors SP600125 or LY294002 drastically impaired cellular motility

(*a-ag respectively). (B) WT explants cultured with MAPK inhibitors and TGFβ1.

Addition of TGFβ1 increased outgrowths to that of untreated in explants treated with

inhibitor UO126 or PD98059 while inhibitors SP600125 or LY294002 significantly

impaired cellular motility, suggesting that TGFβ1 signaling is affected (*a-ab respectively). Results represent mean ± S.E.M. n = 12, *p < 0.05.

216

The addition of TGFβ1 to WT explants increased migration as shown in Figure 5.7C.

Addition of TGFβ1 in conjunction to inhibitor UO126-treated explants increased outgrowth to that of untreated explants (Figure 5.7B). This was similarly observed in inhibitor PD98059-treated explants where no difference in outgrowth was observed between PD98059-treated and untreated explants. In SP600125-treated explants, the

addition of TGFβ1 improved migration and outgrowth was observed at day 4 onwards.

However, no outgrowth was observed in LY294002-treated explants at the end of day 6.

Inhibitor treated Flii+/- explants showed similar results to the WT. The addition of

inhibitor PD98059 showed no difference in migration until day 5 where migration was

significantly decreased compared to the untreated fibroblasts (Figure 5.8A). UO126,

SP600125 and LY294002 all significantly slowed fibroblast outgrowth on day 2 onwards.

SP600125 and LY294002 treated Flii+/- fibroblasts showed almost no outgrowth

indicating that the inhibited pathways are important in cellular migration. At the end of

day 6, it was observed that PD98059-treated explants had reduced fibroblast migration

followed by UO126, SP600125 and LY294002 treated explants. TGFβ1 was shown to

decrease outgrowth migration in Flii+/- fibroblasts (Figure 5.6B). However, TGFβ1

treatment in conjunction with the treatment of inhibitors did not significantly affect

outgrowth of Flii+/- fibroblasts (Figure 5.8B). The outgrowth of Flii+/- treated explants with both inhibitors and TGFβ1 was similar to the outgrowth observed in explants

217

A Flii+/- Explants 300 Untreated *w-af 250 SP600125 LY294002 200 *m-v UO126 PD98059 150 *g-l

100 *a-f Migration Distance Distance Migration (um) 50

0 1 2 3 4 5 6

Days

B Flii+/- Explants

300 Untreated *y-ag 250 SP + b1 LY + b1 200 UO + b1 *p-x PD + b1 150 *g-o 100 *a-f Migration Distance Distance Migration (um) 50

0 1 2 3 4 5 6

Days

218

Figure 5.8

Flii+/- explants treated with MAPK inhibitors and TGFβ1 ligand. Full thickness skin

biopsies were taken from Flii+/-mice. Explants were cultured for a period of 6 days and

cellular outgrowth measured daily. (A) Flii+/- explants treated with MAPK inhibitors only. Explants treated with either inhibitor UO126 or PD98059 had reduced cell migration while inhibitors SP600125 or LY294002 drastically impaired cellular motility

(*a-af respectively). (B) Flii+/- explants cultured with MAPK inhibitors and TGFβ1.

Addition of TGFβ1 showed similar results to A, suggesting that TGFβ1 signaling is not

affected (*a-ag respectively). Results represent mean ± S.E.M. n = 12, *p < 0.05.

219

treated with inhibitors alone. Explants treated with PI3K inhibitor LY294002 showed no

outgrowth.

FliiTg/+ explants treated with inhibitors also showed decreased outgrowths (Figure 5.9A).

UO126 and PD98059 treated explants were observed to have similar outgrowths. FliiTg/+

explants treated with SP600125 and LY294002 inhibitors consistently showed significant

decrease in outgrowth which is similar to WT and Flii+/- explants. Outgrowth was only observed in inhibitor SP600125 treated explants from day 5 onwards. No outgrowth was observed in FliiTg/+ explants treated with inhibitor LY294002. From Figure 5.6D, it is

showed that TGFβ1 increase migration in FliiTg/+outgrowth migration. Similar to WT

explants, FliiTg/+ explants treated with inhibitors UO126 and TGFβ1 also increased

outgrowth migration to be comparable to that of the untreated explants (Figure 5.9B).

Similarly, no differences in outgrowth migration were observed in PD98059 and TGFβ1

treated explants compared to untreated. Conversely, the addition of TGFβ1 to inhibitor

SP600125 treated explants did not increase outgrowth. In fact, no outgrowths were

observed in TGFβ1 treated FliiTg/+ explants in the presence of either SP600125 or

LY294002 inhibitors.

220

A FliiTg/+ Explants

300 Untreated 250 SP600125 LY129004 *y-ag 200 UO126 150 PD98059 *p-x

100 *g-o *a-f

Migration Distance Distance Migration (um) 50

0 1 2 3 4 5 6

Days

B FliiTg/+ Explants

300 Untreated 250 SP + b1 LY + b1 *w-ab 200 UO + b1 150 PD + b1 *p-v

100 *h-o *a-g

Migration Distance Distance Migration (um) 50

0

1 2 3 4 5 6

Days

221

Figure 5.9

FliiTg/+ explants treated with MAPK inhibitors and TGFβ1 ligand. Full thickness skin

biopsies were taken from FliiTg/+ mice. Explants were cultured for a period of 6 days and

cellular outgrowth measured daily. (A) FliiTg/+ explants treated with MAPK inhibitors

only. Explants treated with either inhibitor UO126 or PD98059 had reduced cell

migration while inhibitors SP600125 or LY294002 drastically impaired cellular motility

(*a-ag respectively). (B) FliiTg/+ explants cultured with MAPK inhibitors and TGFβ1.

Addition of TGFβ1 increased outgrowths to that of untreated in explants treated with

inhibitor UO126 or PD98059 while inhibitors SP600125 or LY294002 significantly impair

cellular motility, suggesting that TGFβ1 signaling is affected (*a-ab respectively).

Results represent mean ± S.E.M. n = 12, *p < 0.05.

222

5.3.7 Comparison of WT, Flii+/- and FliiTg/+ outgrowths treated with MAPK inhibitors

The outgrowth migration of WT, Flii+/- and FliiTg/+ explants were compared against one another to determine if there was any difference in outgrowth after treatment with inhibitors and/or TGFβ1. In UO126-treated explants, there were no differences in WT,

Flii+/- and FliiTg/+ outgrowths (Figure 5.10A). However, Flii+/- explants showed a trend of

increased outgrowth migration over WT and FliiTg/+ explants. In PD98059-treated

explants, there were no differences in WT and FliiTg/+ outgrowth migration (Figure

5.10B). Flii+/- explants on the other hand showed significantly increased outgrowth

migration compared to WT and FliiTg/+ on day 4 until day 6. Little outgrowth was

observed in all WT, Flii+/- and FliiTg/+ explants treated with inhibitor SP600125 (Figure

5.10C). Although differences between the groups were observed, the differences were

not significant. This is also the case for WT, Flii+/- and FliiTg/+ explants treated with inhibitor LY294002 (Figure 5.10D). Explants treated with inhibitor LY294002 showed no outgrowth.

The addition of TGFβ1 to inhibitor UO126 treated WT, Flii+/- and FliiTg/+ explants did not

show any differences in outgrowth between them (Figure 5.11A). At day 4, Flii+/-

223

A UO126 B PD98059 200 200 *e-f WT 150 Flii+/- 150 WT *c-d FliiTg/+ Flii+/- *a-b 100 100 *a FliiTg/+ 50 50 Migration Distance Distance Migration (um) 0 Distance Migration (um) 0 1 2 3 4 5 6 1 2 3 4 5 6

Days Days

C SP600125 D LY694002 200 WT 200 150 Flii+/- WT 150 FliiTg/+ Flii+/- 100 *f-g 100 FliiTg/+ *c-e 50 *a-b 50 *a-b *c-e

Distance Migration (um) 0 0 1 2 3 4 5 6 Distance Migration (um) 1 2 3 4 5 6 Days Days

224

Figure 5.10

Comparison of Flii+/-, WT and FliiTg/+ outgrowth in response to MAPK inhibitor

treatments. (A) Explants treated with MEK1/2 inhibitor UO126 showed no significant

differences between the groups. (B) Explants treated with Raf inhibitor PD98059.

Outgrowth migration was further for Flii+/- explants compared to WT and FliiTg/+

explants (*a-f respectively) (C) Explants treated with JNK/SAPK inhibitor SP600125

showed drastically reduced cellular motility. Flii+/- explants had slightly more

outgrowth than WT and FliiTg/+ explants (*a-g respectively) (D) Explants treated with

PI3K inhibitor LY694002 showed drastically reduced outgrowth (*a-e respectively).

Results represent mean ± S.E.M. n = 12, *p < 0.05.

225

A UO126 + TGFβ1 B PD98059 + TGFβ1 200 200 WT WT 150 Flii+/- 150 Flii+/- FliiTg/+ FliiTg/+ 100 *a-b 100

50 50 Migration Distance Distance Migration (um) 0 Distance Migration (um) 0 1 2 3 4 5 6 1 2 3 4 5 6

Days Days

C SP600125 + TGFβ1 D LY694002 + TGFβ1

200 200 WT WT 150 Flii+/- 150 Flii+/- FliiTg/+ FliiTg/+ 100 *e-f 100 *c-d 50 *a-b 50

Migration Distance Distance Migration (um) 0 Distance Migration (um) 0 1 2 3 4 5 6 1 2 3 4 5 6

Days Days

226

Figure 5.11

Comparison of Flii+/-, WT and FliiTg/+ outgrowth in response to MAPK inhibitor and

TGFβ1 ligand treatments. (A) Explants treated with TGFβ1 and MEK1/2 inhibitor

UO126. (B) Explants treated with TGFβ1 and Raf inhibitor PD98059. (C) Explants treated with TGFβ1 and JNK/SAPK inhibitor SP600125 drastically reduced outgrowth in FliiTg/+ explants compared to WT and Flii+/- explants (*a-f respectively) . (D) Explants

treated with TGFβ1 and PI3K inhibitor LY694002 showing no outgrowth in all groups.

Results represent mean ± S.E.M. n = 12, *p < 0.05.

227

explants showed a small increase in outgrowth compared to WT and FliiTg/+ explants but

on day 5 and day 6, no significant difference was observed between the groups. TGFβ1

addition to inhibitor PD98059 treated explants showed no difference in outgrowth in

WT and FliiTg/+ explants (Figure 5.11B). However, both WT and FliiTg/+ outgrowth was

significantly lower than that of Flii+/- explants where it could be clearly observed that

Flii+/- explants treated with inhibitor PD98059 and TGFβ1 had increased outgrowth

migration from day 4 onwards to day 6. In inhibitor SP600125-treated explants in the

presence of TGFβ1, no outgrowth was observed in FliiTg/+ explants (Figure 5.11C). WT and Flii+/- explants had minimal outgrowth and there was no differences in outgrowth

between them. Finally, for explants treated with inhibitor LY294002 and TGFβ1, no

outgrowth was observed in all explant groups (Figure 5.11D). This indicated the

importance of PI3K in cellular migration where the inhibition of this signaling pathway

stops migration even in the presence of TGFβ1.

228

5.4 Discussion

Flii has been implicated as a signaling molecule and is involved in the regulation of

TGFβ expression which consequently affects wound healing. Understanding the

cellular mechanism involved in the process where Flii regulates TGFβ could lead to

therapeutic approaches aimed at improving wound healing and thus scar formation.

TGFβ is an important molecule in the regulation of cellular processes and plays

important roles such as apoptosis, proliferation and migration 185,222,223. Consequently, the action of TGFβs will greatly affect the outcome of wound healing which heavily rely on cellular proliferation and migration, such as keratinocyte migration during re- epithelialization, inflammatory cells influx, fibroblasts migration into the wound matrix and the subsequent production of matrix proteins 111. In previous chapters, the

relationship between Flii and TGFβ has been established, as well as the physical

interactions between Flii and proteins involved in TGFβ signaling. Given that Flii can

interact with Ras proteins 109, implicating Flii in the MAPK signaling pathway which

can also regulate TGFβ signaling, and we set out to study the underlying mechanism by

which Flii regulate TGFβs through the MAPK and PI3K pathways.

The TGFβ signaling pathway requires a well coordinated set of 'activators' such as the

229

binding of cFos and cJun proteins to its promoter and downstream 'effectors' such as the MAPK signaling pathway and Smad proteins 168,201,203. It was first determined whether Flii could affect these proteins by knocking down the expression of the Flii gene. Indeed, cFos and cJun gene expression was decreased in response to the reduction of Flii, suggesting that manipulation of Flii is able to affect the expression of AP-1 proteins, cFos and cJun. As a result, the flow on effects of cFos and cJun expression are likely to impact on the transcriptional activity on the TGFβ promoter to influence TGFβ expression. Smad proteins which signal downstream of TGFβs were also investigated.

Under normal conditions, receptor activated Smad 3 will form complexes with Smad 4 and translocate into the cell nucleus to activate target genes 139. This process is negatively regulated by Smad 7 which prevents the activation of Smad 3 224,225.

Interestingly, the expression of Smad 3 and Smad 4 were found to be down regulated in response to Flii gene reduction. Inhibitory Smad 7 gene expression was however elevated. This implies that cells with reduced Flii could be less responsive to the negative effects of TGFβ signals due to the lack of Smad 3 and Smad 4 proteins and the increase in inhibitory Smad 7.

The effects of TGFβ1 addition to fibroblasts derived from WT, mice heterozygous (Flii+/-) and over-expressing (FliiTg/+) Flii mice was determined. Firstly, the expression of myofibroblasts marker α-smooth muscle actin (α-SMA) was assessed. In unwounded

230

cells, α-SMA expression was found to be highest in FliiTg/+ cells even when TGFβ1 was added. Interestingly, there was no difference in α-SMA expression in any of the groups investigated. As TGFβ1 is known to induce differentiation of fibroblasts into myofibroblasts 226, it was expected that treatment with TGFβ1 would lead to a

significant increase in α-SMA expression in these cells. However, since these cells were

obtained from skin explants, the migrating cells from the explants may have already

differentiated into myofibroblasts therefore mitigating the effects of TGFβ1. In

wounded cells, there were no differences in αSMA expression between WT, Flii+/- and

FliiTg/+ cells suggesting that the fibroblasts have already differentiated into

myofibroblasts. However, the addition of TGFβ1 to Flii+/- fibroblasts increased α-SMA

expression indicating that Flii+/- fibroblasts have less differentiated fibroblasts than WT

and FliiTg/+ cells. This is consistent with the finding that reduction of Flii expression results in a decrease in TGFβ1 expression levels, which may therefore lead to less differentiation of fibroblasts.

TGFβ1 is a known inducer of collagen secretion in fibroblasts 227. In unwounded

fibroblasts, FliiTg/+ fibroblasts secrete significantly more collagen than WT and Flii+/-

fibroblasts. Collagen secretion is increased in response to wounding in WT and FliiTg/+

fibroblasts but Flii+/- fibroblasts collagen secretion remained the same. In cells treated

with TGFβ1, collagen secretion increased significantly with FliiTg/+ fibroblasts secreting

231

the highest levels of collagen. However, collagen secretion in Flii+/- fibroblasts were equivalent to the amount secreted by FliiTg/+ fibroblasts suggesting that Flii+/- fibroblasts

have an internal mechanism preventing collagen production perhaps by suppressing

TGFβ1 expression but adding exogenous TGFβ1 overrides that mechanism therefore

increasing collagen levels in Flii+/- fibroblasts. This may also explain the impairment of

wound healing in FliiTg/+ mice 111 as excessive collagen production in FliiTg/+ fibroblasts

lead to fibrosis which contributes to impaired wound healing 228.

Cellular proliferation is key to the wound healing process. Proliferation in primary

fibroblasts was determined following the addition of MAPK inhibitors and TGFβ1.

Increased proliferation of Flii+/- fibroblasts was consistent with previous results showing

highest proliferation in Flii+/- fibroblasts followed by WT and FliiTg/+ fibroblasts 111. A

dose response concentration of 500pg/ml, 1000pg/ml and 2000pg/ml of TGFβ1 were

added to the cells representing low, physiological and high concentrations respectively.

Proliferation in WT and Flii+/- fibroblasts was increased in response to TGFβ1 addition

in all concentrations indicating the sensitivity of cellular proliferation to TGFβ1. FliiTg/+

fibroblasts proliferation on the other hand was decreased. This experiment showed that

addition of TGFβ1 have different effects in cells expressing different levels of Flii.

232

MAPK pathway inhibitors were used to investigate the underlying mechanism of TGFβ

regulation by Flii. It is noted that the non-specific nature of the MAPK inhibitors used in this study are taken into consideration such that these inhibitors do not specifically target a single signaling pathway. It is more likely that MAPK inhibitors such as

SP600125 can inhibit other kinase activities, which results in different cellular responses

229. In addition, there are a huge number of kinases that exists and these inhibitors have

only been tested for a small percentage of them. Therefore, there is a need to confirm the data using different approaches such as using different inhibitors targeting the same kinase activity and using more specific methods including targeted knockdown of the kinase.

As the MAPK signaling pathway regulates cellular proliferation 153, there is no doubt

that proliferation in WT, Flii+/- and FliiTg/+ fibroblasts will be affected. The addition of

MAPK inhibitors (UO126, PD98059, SP600125 and LY294002) to WT fibroblasts resulted

in an increase in proliferation opposite to FliiTg/+ fibroblasts where reduced proliferation

was observed. This is interesting as it was expected that the addition of inhibitors

would decrease proliferation. However, it could be due to the multiple signaling cross-

talks within the MAPK signaling pathway 230. As a result, other signaling pathways may compensate for the loss of one signaling column. Proliferation in Flii+/- fibroblasts were significantly increased in all inhibitor treatments except for fibroblasts treated with the

233

JNK/SAPK inhibitor SP600125. In fact, only the addition of this inhibitor prevented the increase in proliferation characteristic of Flii+/- fibroblasts indicating that Flii+/- fibroblasts may be dependent on the activity of the JNK/SAPK signaling pathway for its effects on proliferation to occur.

TGFβ1 addition to primary explants were also found to have different effects on WT,

Flii+/- and FliiTg/+ outgrowth migration. In normal untreated explants, Flii+/- explants had the greatest outgrowth response followed by WT and FliiTg/+ explants, which is consistent with the proliferation results and study showing improved migration in Flii knockdown cells 111. The effects of TGFβ1 on cellular migration is still a debatable topic where it has been shown to either increase or decrease cellular migration 185,186,231,232. WT and FliiTg/+ explants treated with TGFβ1 showed increased outgrowth but Flii+/- explants showed decreased outgrowth. With similarity to the proliferation result, the addition of

TGFβ1 appeared to have different effects on cells expressing differential Flii levels.

Having established the effects of TGFβ1 addition on WT, Flii+/- and FliiTg/+ outgrowth migration, we now investigated the effects of MAPK inhibitors on WT, Flii+/- and FliiTg/+ outgrowth migration. The addition of MAPK kinase inhibitor to explants decrease outgrowth in all WT, Flii+/- and FliiTg/+ explants which shows that the MAPK signaling pathway plays an important role in cellular migration 154. Inhibitor PD98059 and UO126 which inhibits Raf and MEK 1/2 respectively affected migration to a lesser degree than

234

JNK/SAPK inhibitor SP600125 or PI3K inhibitor LY294002. These results indicated that

cellular migration can be more dependent on a MAPK signaling column such as the

activation of JNK/SAPK or PI3K signaling pathways, without which would lead to

significant loss of motility. On the other hand, MAPK signaling columns such as ERK

are also important but can be compensated by other signaling pathways via the

accessibility of multiple cross-talks present in the MAPK signaling pathway 233.

The addition of TGFβ1 to MAPK inhibitor treated explants allowed us to investigate whether TGFβ1 would still have an effect on cellular outgrowth when the MAPK signaling pathway is inhibited. In all WT, Flii+/- and FliiTg/+ explants, the treatment with

inhibitors SP600125 or LY294002 results in negligible outgrowth even when treated

with TGFβ1. This shows that TGFβ1 by itself is not sufficient to affect cellular

outgrowth; it also requires the cooperation of MAPK signaling pathways in order to

have an effect. Therefore, it is possible that the feedback loop of TGFβ1 activating

MAPK pathway may also contribute to its activity and efficacy 234. This observation is

consistently seen in all WT, Flii+/- and FliiTg/+ explants treated with inhibitors and TGFβ1.

Comparing WT, Flii+/- and FliiTg/+ explants treated with UO126 did not yield any differences between the groups. UO126 inhibits the phosphorylation of ERK1/2 from

MEK 158. As ERK is known to be involved in cellular migration 235 and Flii may be acting

through this pathway, the inhibition of this ERK will have an impact on the

235

downstream processes of Flii. Therefore, this may be the reason why there are no

differences between WT, Flii+/- and FliiTg/+ outgrowth migration. This is also similar to

the addition of Raf inhibitor PD98059. However, the outgrowth of cells in Flii+/- explants

went further than WT and FliiTg/+ explants. This suggests that although the inhibition of

MEK by Raf causes lesser outgrowth, compensation from other signaling cross-talks can

offset the decreased outgrowth. This indicated that increased in outgrowth migration of

Flii+/- explants is not heavily dependent on Raf activity.

In summary, we have shown that manipulation of Flii expression also affects genes

important for TGFβ expressions and activities. We have also shown that MAPK

signaling pathway is important in cellular proliferation and migration and inhibition of

any pathway within the MAPKs affects both cellular processes. Addition of PI3K

inhibitor LY294002 and JNK/SAPK inhibitor SP600125 results in severe outgrowth

inhibition. This indicated that these two pathways are essential in allowing cellular

migration to proceed. The treatment of Raf inhibitor PD98059 or Mek 1/2 inhibitor

UO126 also affected outgrowth migration but to a lesser degree. As cross-talks are prevalent in these signaling pathways, the inhibition of these pathways may be compensated for by other MAPK signaling cascades 230,233. The similarities between WT,

Flii+/- and FliiTg/+ groups treated with inhibitors could be explained by the activity of Flii.

Its effect is not mediated through just one specific pathway. It requires the cooperation

236

between MAPK signaling pathways to affect downstream processes such as migration.

Therefore the inhibition of a specific pathway results in little or no significant differences between the groups. In conclusion, the regulation of TGFβ by Flii is most likely mediated through the MAPK signaling pathway.

237

CHAPTER SIX

General discussion

6.1 Discussion

Wound healing is orchestrated by a complex interplay between different cells including

inflammatory cells, keratinocytes and fibroblasts, which are governed by a cocktail of

signaling molecules such as hormones and cytokines. In addition, the actin cytoskeleton

is the underlying structure integral to the wound healing process as cells require the

actin cytoskeleton to migrate by forming protusions such as lamellipodias and

filopodias. The actin cytoskeleton is regulated by the gelsolin family of actin remodeling

proteins. Although the gelsolin family has been identified as the chief regulator of actin,

this thesis has identified additional roles for members of the gelsolin family involved in

wound healing. This thesis aims to examine the roles of two proteins in the gelsolin family of actin binding proteins, gelsolin and Flii in the process of wound healing. The insignificant outcome of the gelsolin study meant that the thesis is split into two parts with more emphasis on the latter.

6.1.1 Gelsolin and the AR

Several studies have identified gelsolin as a AR co-activator 81,89,178,236, which lead to the

prediction that gelsolin is of importance in the androgen mediated wound healing

238

process. This study has looked at possible roles of gelsolin in the androgen mediated wound repair by using gelsolin siRNA knock down fibroblasts and the addition of hormone, DHT. The knockdown of gelsolin in HFFs resulted in decreased proliferation and impaired migration. However, the addition of hormone DHT reversed the effect and cellular proliferation and migration were increased. When this hypothesize was subsequently shown to be inaccurate, the alternative explanation was that even though gelsolin is an co-activator of AR, it is not directly involved in DHT stimulated cellular proliferation and migration. A second possibility is that functional redundancy may have occurred. This implies that the co-activating function of gelsolin can be compensated by the availability of other co-activators including members of the gelsolin family such as supervillin which is also a co-activator for AR 177. However, it is unknown whether gelsolin localization into the nucleus following hormone stimulation have any additional roles as it seems unlikely that nuclear gelsolin has redundant functions in evolutionary advance mammals. In addition, the importance of gelsolin interaction with AR may not lie in the androgen mediated wound healing process but in other cellular processes. It is also interesting to see whether gelsolin is also a co- activator for other signaling proteins. This may be of interest for future research to fully understand the complete function of the gelsolin family.

239

6.1.2 Flii and TGFβs

Following studies reporting the negative effects of TGFβ1 and the beneficial effects of

TGFβ3 116,117,192, chapter 3 aims to identify the relationship between Flii and the TGFβs by using an in vivo wound healing model and detecting TGFβ isoforms expression throughout the wound healing time-course. Here, we found that Flii deficiency mice correlates with low to high ratios of TGFβ1 to TGFβ3 expression. Conversely, in Flii over-expressing mice correlates with high to low ratios of TGFβ1 to TGFβ3. To further substantiate the finding that Flii deficiency results in low TGFβ1 to high TGFβ3 expression ratios, we used Flii siRNA gene silencing technology to knockdown Flii gene expression. Consistent to the in vivo wound healing results, reducing Flii gene expression is also correlated with lower levels of TGFβ1 and higher levels of TGFβ3.

This showed that manipulation of Flii could lead to different expression of TGFβ isoforms. Given the beneficial effects of TGFβ3, this could explain the improved healing in mice heterozygous for Flii. The opposite could also be said for mice over-expressing

Flii. As the relationship between Flii and TGFβ has been established, it will be interesting to determine whether the manipulation of Flii in TGFβ null mice will have any differences in the outcome of wound healing.

In chapter 4 and chapter 5, the underlying mechanisms for Flii regulation of TGFβ were

240

investigated. As proteins with a LRR domain interacts with other proteins, it is

therefore likely that Flii will interact with proteins given that Flii also has a LRR

domain. It was revealed that Flii directly associates with multiple proteins involved in

TGFβ expression and signaling as well as TGFβ itself. Flii associates with c-Fos and c-

Jun proteins which are involved in the regulation of TGFβ expression. In addition, it

was also found that Flii associates with Akt and Smad proteins involved in downstream

TGFβ signaling. These results indicated a potential mechanistic link suggesting that Flii could potentially regulate TGFβs.

Here, two possible roles that Flii can act as a regulator for wound healing are proposed.

Firstly, Flii may function as a co-activator for target gene expression. Since co-activating

functions for Flii been previously described 105, it is therefore possible that the

interactions of Flii with transcriptional activators such as Smad 2/3 or cFos and cJun dimers, have co-activating roles. These interactions could represent the formation of a

co-activator complex, which is part of a larger transcriptional activating complex that

binds to DNA and regulate the expression of target genes (Figure 6.1). The second

possible role of Flii involves the sequestering of the target protein, which affects

downstream signaling activities. For example, B23 is a protein that play important roles

in cellular proliferation and cell death 237. The binding of Akt to B23 stabilizes B23 and

prevent it from degradation by caspases. In this case, Flii can complete with B23 for Akt

241

Stimulant Flii Akt

Protein Activities

Flii Flii TGFβ

Flii Flii 4 2/3 2/3 Transcription Fos/Jun Transcription X Factors Dimers X Factors Smad Smad Smad

Gene Expression Gene Expression

Figure 6.1

Theoretical representation of Flii interactions and functions.

242

which affects the outcome of cellular proliferation or cell death. As a result, these interactions may account for the differences in wound healing observed in WT, Flii+/- and FliiTg/+ mice.

In addition to direct interaction, manipulation of Flii can also affect the gene expression of cFos, cJun, Akt and Smad 2/3 proteins. Since TGFβ can be regulated by the flow through effects of the MAPK signaling pathway and that Flii has links to the MAPK pathway, there is a high possibility that Flii may regulate TGFβ expression through the

MAPK signaling pathway. We first investigated the effects of TGFβ1 addition to primary fibroblasts derived from Flii heterozygous and over-expressing mice. It was found that the addition of TGFβ1 increases proliferation in Flii+/- fibroblasts but decreases outgrowth migration and vice versa in FliiTg/+ fibroblasts. This study showed that cellular responses to TGFβ1 ligand addition differ between WT, Flii+/- and FliiTg/+ fibroblasts. Treatment with MAPK inhibitors, UO126 (Mek1/2 inhibitor), PD98059 (Raf inhibitor), SP600125 (JNK/SAPK inhibitor) and LY294002 (PI3K inhibitor) did not yield any significant migratory differences between WT, Flii+/- and FliiTg/+ groups. This is also similarly observed if TGFβ1 ligand was added to the inhibitor treated cells. However, fibroblasts treated with MAPK inhibitors showed decreased proliferation in Flii+/- but increased proliferation in FliiTg/+ cells. The results from chapter 5 suggest that TGFβ regulation by Flii is mediated through the MAPK signaling pathway. It is also

243

interesting to speculate that since Flii can directly interact as well as influence the expression of cFos, cJun, Akt and Smad 2/3 proteins, Flii could therefore have direct and indirect roles in the regulation of TGFβ. Taken together, this thesis has reveal several novel findings and has contributed to the better understanding of Flii involvement in the wound healing process. This will contribute to a better therapeutic design by which wound healing might be improved.

6.1.3 Future Directions

At time of writing, a number of interacting proteins have now been identified for Flii including those revealed in this thesis. It would be worthwhile to determine precisely where the regions of interactions are located. This can be done using various Flii point mutants and GST-pulldown techniques. As Flii is regarded as a negative regulator for wound healing 110,111, the interactions of Flii with proteins identified in this chapter may be of a repressive nature. Therefore, by determining the significance of interactions on

Flii, it will provide an insight into the roles played by Flii such that if point mutation in

Flii results in the specific disruption of one individual interaction and not the others.

The analysis of changes in gene expression in response to Flii depletion indicated a role of Flii in the cell nucleus and possibly have roles in the MAPK signaling pathway given that Flii has links to the pathway via its interaction with Ras proteins 109. It will be

244

interesting to investigate this possibility in future studies using MAPK pathway inhibitors and in vitro reporter assays. Another point to consider is the specificity of the inhibitors used in this study. Since current MAPK inhibitors can affect unknown kinase activities, other methods including the use of more specific approach such as targeted knockdown of the kinase should be employed to confirm the results from this study.

These studies will contribute to the confidence on how Flii is working through the various signaling pathways. The establishment of a role for Flii in the MAPK signaling pathway will open new doors for research into the function of Flii in many aspects of cellular processes other than wound healing.

245

BIBLIOGRAPHY

1 Baum CL, Arpey CJ. Normal cutaneous wound healing: clinical correlation with

cellular and molecular events. Dermatol Surg 2005; 31: 674-86; discussion 86.

2 Mosesson MW, Siebenlist KR, Meh DA. The structure and biological features of

fibrinogen and fibrin. Annals of the New York Academy of Sciences 2001; 936: 11-30.

3 Werner S, Grose R. Regulation of wound healing by growth factors and

cytokines. Physiological reviews 2003; 83: 835-70.

4 Monaco JL, Lawrence WT. Acute wound healing an overview. Clinics in plastic

surgery 2003; 30: 1-12.

5 Singer AJ, Clark RA. Cutaneous wound healing. The New England journal of

medicine 1999; 341: 738-46.

6 Santucci M, Borgognoni L, Reali UM et al. Keloids and hypertrophic scars of

Caucasians show distinctive morphologic and immunophenotypic profiles.

Virchows Arch 2001; 438: 457-63.

7 Wolfram D, Tzankov A, Pulzl P et al. Hypertrophic scars and keloids--a review of

their pathophysiology, risk factors, and therapeutic management. Dermatol Surg

2009; 35: 171-81.

246

8 Gurtner GC, Werner S, Barrandon Y et al. Wound repair and regeneration. Nature

2008; 453: 314-21.

9 Hebda PA, Collins MA, Tharp MD. Mast cell and myofibroblast in wound

healing. Dermatologic clinics 1993; 11: 685-96.

10 Lawrence WT. Physiology of the acute wound. Clinics in plastic surgery 1998; 25:

321-40.

11 Hubner G, Brauchle M, Smola H et al. Differential regulation of pro-

inflammatory cytokines during wound healing in normal and glucocorticoid-

treated mice. Cytokine 1996; 8: 548-56.

12 Broughton G, 2nd, Janis JE, Attinger CE. Wound healing: an overview. Plastic and

reconstructive surgery 2006; 117: 1e-S-32e-S.

13 Lee MY, Ehrlich HP. Influence of vanadate on migrating fibroblast orientation

within a fibrin matrix. Journal of cellular physiology 2008; 217: 72-6.

14 Greiling D, Clark RA. Fibronectin provides a conduit for fibroblast

transmigration from collagenous stroma into fibrin clot provisional matrix.

Journal of cell science 1997; 110 ( Pt 7): 861-70.

15 Abraham DJ, Shiwen X, Black CM et al. Tumor necrosis factor alpha suppresses

the induction of connective tissue growth factor by transforming growth factor-

beta in normal and scleroderma fibroblasts. The Journal of biological chemistry 2000;

275: 15220-5.

247

16 Canty EG, Kadler KE. Procollagen trafficking, processing and fibrillogenesis.

Journal of cell science 2005; 118: 1341-53.

17 Desmouliere A, Chaponnier C, Gabbiani G. Tissue repair, contraction, and the

myofibroblast. Wound Repair Regen 2005; 13: 7-12.

18 Li J, Zhang YP, Kirsner RS. Angiogenesis in wound repair: angiogenic growth

factors and the extracellular matrix. Microscopy research and technique 2003; 60:

107-14.

19 Ehrlich HP, Krummel TM. Regulation of wound healing from a connective tissue

perspective. Wound Repair Regen 1996; 4: 203-10.

20 Werb Z, Tremble P, Damsky CH. Regulation of extracellular matrix degradation

by cell-extracellular matrix interactions. Cell Differ Dev 1990; 32: 299-306.

21 Amadeu TP, Coulomb B, Desmouliere A et al. Cutaneous wound healing:

myofibroblastic differentiation and in vitro models. The international journal of

lower extremity wounds 2003; 2: 60-8.

22 Cairns NJ, Lee VM, Trojanowski JQ. The cytoskeleton in neurodegenerative

diseases. The Journal of pathology 2004; 204: 438-49.

23 Oshima RG. Intermediate filaments: a historical perspective. Experimental cell

research 2007; 313: 1981-94.

24 Monton H, Nogues C, Rossinyol E et al. QDs versus Alexa: reality of promising

tools for immunocytochemistry. Journal of 2009; 7: 4.

248

25 Brangwynne CP, MacKintosh FC, Kumar S et al. Microtubules can bear enhanced

compressive loads in living cells because of lateral reinforcement. The Journal of

2006; 173: 733-41.

26 Esue O, Carson AA, Tseng Y et al. A direct interaction between actin and

filaments mediated by the tail domain of vimentin. The Journal of

biological chemistry 2006; 281: 30393-9.

27 Valiron O, Caudron N, Job D. Microtubule dynamics. Cell Mol Life Sci 2001; 58:

2069-84.

28 Holzbaur EL. Motor neurons rely on motor proteins. Trends in cell biology 2004;

14: 233-40.

29 Goldman RD, Grin B, Mendez MG et al. Intermediate filaments: versatile

building blocks of cell structure. Current opinion in cell biology 2008; 20: 28-34.

30 Parry DA, Strelkov SV, Burkhard P et al. Towards a molecular description of

intermediate filament structure and assembly. Experimental cell research 2007; 313:

2204-16.

31 Capetanaki Y, Bloch RJ, Kouloumenta A et al. Muscle intermediate filaments and

their links to membranes and membranous organelles. Experimental cell research

2007; 313: 2063-76.

32 Green KJ, Simpson CL. Desmosomes: new perspectives on a classic. The Journal of

investigative dermatology 2007; 127: 2499-515.

249

33 Feuer G, Molnar F, et al. Studies on the composition and polymerization of actin.

Hungarica acta physiologica 1948; 1: 150-63.

34 Sheterline P, Clayton J, Sparrow J. Actin. Protein profile 1995; 2: 1-103.

35 Lambrechts A, Van Troys M, Ampe C. The actin cytoskeleton in normal and

pathological cell motility. The international journal of & cell biology

2004; 36: 1890-909.

36 Steinmetz MO, Goldie KN, Aebi U. A correlative analysis of actin filament

assembly, structure, and dynamics. The Journal of cell biology 1997; 138: 559-74.

37 Naumanen P, Lappalainen P, Hotulainen P. Mechanisms of actin stress fibre

assembly. Journal of microscopy 2008; 231: 446-54.

38 Goffin JM, Pittet P, Csucs G et al. Focal adhesion size controls tension-dependent

recruitment of alpha-smooth muscle actin to stress fibers. The Journal of cell

biology 2006; 172: 259-68.

39 Small JV, Rottner K, Kaverina I et al. Assembling an actin cytoskeleton for cell

attachment and movement. Biochimica et biophysica acta 1998; 1404: 271-81.

40 Hotulainen P, Lappalainen P. Stress fibers are generated by two distinct actin

assembly mechanisms in motile cells. The Journal of cell biology 2006; 173: 383-94.

41 Small JV, Stradal T, Vignal E et al. The lamellipodium: where motility begins.

Trends in cell biology 2002; 12: 112-20.

250

42 Jaffe AB, Hall A. Rho GTPases: biochemistry and biology. Annual review of cell

and developmental biology 2005; 21: 247-69.

43 Pollard TD. Regulation of actin filament assembly by Arp2/3 complex and

formins. Annual review of and biomolecular structure 2007; 36: 451-77.

44 Mattila PK, Lappalainen P. Filopodia: molecular architecture and cellular

functions. Nature reviews 2008; 9: 446-54.

45 Hu K, Ji L, Applegate KT et al. Differential transmission of actin motion within

focal adhesions. Science (New York, N.Y 2007; 315: 111-5.

46 Delon I, Brown NH. Integrins and the actin cytoskeleton. Current opinion in cell

biology 2007; 19: 43-50.

47 Laukaitis CM, Webb DJ, Donais K et al. Differential dynamics of alpha 5 integrin,

paxillin, and alpha-actinin during formation and disassembly of adhesions in

migrating cells. The Journal of cell biology 2001; 153: 1427-40.

48 Rottner K, Hall A, Small JV. Interplay between Rac and Rho in the control of

substrate contact dynamics. Curr Biol 1999; 9: 640-8.

49 Zaidel-Bar R, Ballestrem C, Kam Z et al. Early molecular events in the assembly

of matrix adhesions at the leading edge of migrating cells. Journal of cell science

2003; 116: 4605-13.

50 Le Clainche C, Carlier MF. Regulation of actin assembly associated with

protrusion and adhesion in cell migration. Physiological reviews 2008; 88: 489-513.

251

51 Bryan J. Gelsolin has three actin-binding sites. The Journal of cell biology 1988; 106:

1553-62.

52 Sun HQ, Yamamoto M, Mejillano M et al. Gelsolin, a multifunctional actin

regulatory protein. The Journal of biological chemistry 1999; 274: 33179-82.

53 Archer SK, Claudianos C, Campbell HD. Evolution of the gelsolin family of

actin-binding proteins as novel transcriptional coactivators. Bioessays 2005; 27:

388-96.

54 Campbell HD, Fountain S, Young IG et al. Genomic structure, evolution, and

expression of human FLII, a gelsolin and leucine-rich-repeat family member:

overlap with LLGL. Genomics 1997; 42: 46-54.

55 Arai M, Kwiatkowski DJ. Differential developmentally regulated expression of

gelsolin family members in the mouse. Dev Dyn 1999; 215: 297-307.

56 Witke W, Li W, Kwiatkowski DJ et al. Comparisons of CapG and gelsolin-null

macrophages: demonstration of a unique role for CapG in receptor-mediated

ruffling, phagocytosis, and vesicle rocketing. The Journal of cell biology 2001; 154:

775-84.

57 Yin HL. Gelsolin: calcium- and polyphosphoinositide-regulated actin-

modulating protein. Bioessays 1987; 7: 176-9.

58 Urosev D, Ma Q, Tan AL et al. The structure of gelsolin bound to ATP. Journal of

molecular biology 2006; 357: 765-72.

252

59 Yamamoto H, Ito H, Nakamura H et al. Human plasma gelsolin binds adenosine

triphosphate. Journal of biochemistry 1990; 108: 505-6.

60 Tanskanen M, Paetau A, Salonen O et al. Retraction. Withdrawn: severe ataxia

with neuropathy in hereditary gelsolin amyloidosis. 2009; 16: 246.

61 Tanskanen M, Paetau A, Salonen O et al. Severe ataxia with neuropathy in

hereditary gelsolin amyloidosis. Amyloid 2009: 1-7.

62 Carrwik C, Stenevi U. Lattice corneal dystrophy, gelsolin type (Meretoja's

syndrome). Acta Ophthalmol 2009.

63 Qiao H, Koya RC, Nakagawa K et al. Inhibition of Alzheimer's amyloid-beta

peptide-induced reduction of mitochondrial and

neurotoxicity by gelsolin. Neurobiol Aging 2005; 26: 849-55.

64 Ji L, Chauhan A, Wegiel J et al. Gelsolin is proteolytically cleaved in the brains of

individuals with Alzheimer's disease. J Alzheimers Dis 2009; 18: 105-11.

65 Chauhan V, Ji L, Chauhan A. Anti-amyloidogenic, anti-oxidant and anti-

apoptotic role of gelsolin in Alzheimer's disease. Biogerontology 2008; 9: 381-9.

66 Thor AD, Edgerton SM, Liu S et al. Gelsolin as a negative prognostic factor and

effector of motility in erbB-2-positive epidermal growth factor receptor-positive

breast cancers. Clin Cancer Res 2001; 7: 2415-24.

253

67 Sagawa N, Fujita H, Banno Y et al. Gelsolin suppresses tumorigenicity through

inhibiting PKC activation in a human lung cancer cell line, PC10. British journal of

cancer 2003; 88: 606-12.

68 Geng YJ, Azuma T, Tang JX et al. Caspase-3-induced gelsolin fragmentation

contributes to actin cytoskeletal collapse, nucleolysis, and apoptosis of vascular

smooth muscle cells exposed to proinflammatory cytokines. European journal of

cell biology 1998; 77: 294-302.

69 Azuma T, Witke W, Stossel TP et al. Gelsolin is a downstream effector of for

fibroblast motility. The EMBO journal 1998; 17: 1362-70.

70 Kamada S, Kusano H, Fujita H et al. A cloning method for caspase substrates that

uses the yeast two-hybrid system: cloning of the antiapoptotic gene gelsolin.

Proceedings of the National Academy of Sciences of the United States of America 1998;

95: 8532-7.

71 Kothakota S, Azuma T, Reinhard C et al. Caspase-3-generated fragment of

gelsolin: effector of morphological change in apoptosis. Science (New York, N.Y

1997; 278: 294-8.

72 Dong Y, Asch HL, Medina D et al. Concurrent deregulation of gelsolin and cyclin

D1 in the majority of human and rodent breast cancers. Int J Cancer 1999; 81: 930-

8.

254

73 Asch HL, Winston JS, Edge SB et al. Down-regulation of gelsolin expression in

human breast ductal carcinoma in situ with and without invasion. Breast Cancer

Res Treat 1999; 55: 179-88.

74 Lee HK, Driscoll D, Asch H et al. Downregulated gelsolin expression in

hyperplastic and neoplastic lesions of the prostate. Prostate 1999; 40: 14-9.

75 Ohtsu M, Sakai N, Fujita H et al. Inhibition of apoptosis by the actin-regulatory

protein gelsolin. The EMBO journal 1997; 16: 4650-6.

76 Fujita H, Laham LE, Janmey PA et al. Functions of [His321]gelsolin isolated from

a flat revertant of ras-transformed cells. European journal of biochemistry / FEBS

1995; 229: 615-20.

77 Azuma T, Koths K, Flanagan L et al. Gelsolin in complex with

phosphatidylinositol 4,5-bisphosphate inhibits caspase-3 and -9 to retard

apoptotic progression. The Journal of biological chemistry 2000; 275: 3761-6.

78 Witke W, Sharpe AH, Hartwig JH et al. Hemostatic, inflammatory, and fibroblast

responses are blunted in mice lacking gelsolin. Cell 1995; 81: 41-51.

79 Lu M, Witke W, Kwiatkowski DJ et al. Delayed retraction of filopodia in gelsolin

null mice. The Journal of cell biology 1997; 138: 1279-87.

80 Serrander L, Skarman P, Rasmussen B et al. Selective inhibition of IgG-mediated

phagocytosis in gelsolin-deficient murine neutrophils. J Immunol 2000; 165: 2451-

7.

255

81 Nishimura K, Ting HJ, Harada Y et al. Modulation of androgen receptor

transactivation by gelsolin: a newly identified androgen receptor coregulator.

Cancer research 2003; 63: 4888-94.

82 Archer SK, Behm CA, Claudianos C et al. The flightless I protein and the gelsolin

family in nuclear hormone receptor-mediated signalling. Biochemical Society

transactions 2004; 32: 940-2.

83 Ashcroft GS, Mills SJ, Ashworth JJ. Ageing and wound healing. Biogerontology

2002; 3: 337-45.

84 Labrie F, Luu-The V, Labrie C et al. DHEA and its transformation into androgens

and estrogens in peripheral target tissues: intracrinology. Frontiers in

neuroendocrinology 2001; 22: 185-212.

85 Gilliver SC, Ashworth JJ, Ashcroft GS. The hormonal regulation of cutaneous

wound healing. Clinics in dermatology 2007; 25: 56-62.

86 Kanda N, Watanabe S. Regulatory roles of sex hormones in cutaneous biology

and immunology. Journal of dermatological science 2005; 38: 1-7.

87 Ashcroft GS, Greenwell-Wild T, Horan MA et al. Topical estrogen accelerates

cutaneous wound healing in aged humans associated with an altered

inflammatory response. The American journal of pathology 1999; 155: 1137-46.

256

88 Nitsch SM, Wittmann F, Angele P et al. Physiological levels of 5 alpha-

dihydrotestosterone depress wound immune function and impair wound

healing following trauma-hemorrhage. Arch Surg 2004; 139: 157-63.

89 Ashcroft GS, Mills SJ. Androgen receptor-mediated inhibition of cutaneous

wound healing. The Journal of clinical investigation 2002; 110: 615-24.

90 Gilliver SC, Ashworth JJ, Mills SJ et al. Androgens modulate the inflammatory

response during acute wound healing. Journal of cell science 2006; 119: 722-32.

91 Fimmel S, Zouboulis CC. Influence of physiological androgen levels on wound

healing and immune status in men. Aging Male 2005; 8: 166-74.

92 Hermanson O, Glass CK, Rosenfeld MG. Nuclear receptor coregulators: multiple

modes of modification. Trends in endocrinology and metabolism: TEM 2002; 13: 55-

60.

93 Campbell HD, Schimansky T, Claudianos C et al. The

flightless-I gene involved in gastrulation and muscle degeneration encodes

gelsolin-like and leucine-rich repeat domains and is conserved in Caenorhabditis

elegans and humans. Proceedings of the National Academy of Sciences of the United

States of America 1993; 90: 11386-90.

94 Davy DA, Ball EE, Matthaei KI et al. The flightless I protein localizes to actin-

based structures during embryonic development. Immunol Cell Biol 2000; 78: 423-

9.

257

95 Miklos GL, De Couet HG. The mutations previously designated as flightless-I3,

flightless-O2 and standby are members of the W-2 lethal complementation group

at the base of the X-chromosome of Drosophila melanogaster. J Neurogenet 1990;

6: 133-51.

96 Straub KL, Stella MC, Leptin M. The gelsolin-related flightless I protein is

required for actin distribution during cellularisation in Drosophila. Journal of cell

science 1996; 109 ( Pt 1): 263-70.

97 Chen KS, Gunaratne PH, Hoheisel JD et al. The human homologue of the

Drosophila melanogaster flightless-I gene (flil) maps within the Smith-Magenis

microdeletion critical region in 17p11.2. Am J Hum Genet 1995; 56: 175-82.

98 Campbell HD, Fountain S, McLennan IS et al. Fliih, a gelsolin-related cytoskeletal

regulator essential for early mammalian embryonic development. Molecular and

cellular biology 2002; 22: 3518-26.

99 Deng H, Xia D, Fang B et al. The Flightless I homolog, fli-1, regulates

anterior/posterior polarity, asymmetric cell division and ovulation during

Caenorhabditis elegans development. Genetics 2007; 177: 847-60.

100 Goshima M, Kariya K, Yamawaki-Kataoka Y et al. Characterization of a novel

Ras-binding protein Ce-FLI-1 comprising leucine-rich repeats and gelsolin-like

domains. Biochemical and biophysical research communications 1999; 257: 111-6.

258

101 Bella J, Hindle KL, McEwan PA et al. The leucine-rich repeat structure. Cell Mol

Life Sci 2008; 65: 2307-33.

102 Kobe B, Kajava AV. The leucine-rich repeat as a protein recognition motif.

Current opinion in structural biology 2001; 11: 725-32.

103 Liu YT, Yin HL. Identification of the binding partners for flightless I, A novel

protein bridging the leucine-rich repeat and the gelsolin superfamilies. The

Journal of biological chemistry 1998; 273: 7920-7.

104 Seward ME, Easley CAt, McLeod JJ et al. Flightless-I, a gelsolin family member

and transcriptional regulator, preferentially binds directly to activated cytosolic

CaMK-II. FEBS letters 2008; 582: 2489-95.

105 Lee YH, Campbell HD, Stallcup MR. Developmentally essential protein flightless

I is a nuclear receptor coactivator with actin binding activity. Molecular and

cellular biology 2004; 24: 2103-17.

106 Jeong KW, Lee YH, Stallcup MR. Recruitment of the SWI/SNF chromatin-

remodeling complex to steroid hormone-regulated promoters by nuclear

receptor coactivator Flightless-I. The Journal of biological chemistry 2009.

107 Xu J, Liao L, Qin J et al. Identification of Flightless-I as a substrate of the cytokine-

independent survival kinase CISK. The Journal of biological chemistry 2009; 284:

14377-85.

259

108 Li J, Yin HL, Yuan J. Flightless-I regulates proinflammatory caspases by

selectively modulating intracellular localization and caspase activity. The Journal

of cell biology 2008; 181: 321-33.

109 Davy DA, Campbell HD, Fountain S et al. The flightless I protein colocalizes with

actin- and microtubule-based structures in motile Swiss 3T3 fibroblasts: evidence

for the involvement of PI 3-kinase and Ras-related small GTPases. Journal of cell

science 2001; 114: 549-62.

110 Adams DH, Ruzehaji N, Strudwick XL et al. Attenuation of Flightless I, an actin-

remodelling protein, improves burn injury repair via modulation of transforming

growth factor (TGF)-beta1 and TGF-beta3. Br J Dermatol 2009; 161: 326-36.

111 Cowin AJ, Adams DH, Strudwick XL et al. Flightless I deficiency enhances

wound repair by increasing cell migration and proliferation. The Journal of

pathology 2007; 211: 572-81.

112 Kopecki Z, Luchetti MM, Adams DH et al. Collagen loss and impaired wound

healing is associated with c-Myb deficiency. The Journal of pathology 2007; 211:

351-61.

113 Chin D, Boyle GM, Parsons PG et al. What is transforming growth factor-beta

(TGF-beta)? British journal of plastic surgery 2004; 57: 215-21.

260

114 Frank S, Madlener M, Werner S. Transforming growth factors beta1, beta2, and

beta3 and their receptors are differentially regulated during normal and

impaired wound healing. The Journal of biological chemistry 1996; 271: 10188-93.

115 Hsu M, Peled ZM, Chin GS et al. Ontogeny of expression of transforming growth

factor-beta 1 (TGF-beta 1), TGF-beta 3, and TGF-beta receptors I and II in fetal rat

fibroblasts and skin. Plastic and reconstructive surgery 2001; 107: 1787-94;

discussion 95-6.

116 Shah M, Foreman DM, Ferguson MW. Neutralisation of TGF-beta 1 and TGF-

beta 2 or exogenous addition of TGF-beta 3 to cutaneous rat wounds reduces

scarring. Journal of cell science 1995; 108 ( Pt 3): 985-1002.

117 Shah M, Foreman DM, Ferguson MW. Neutralising antibody to TGF-beta 1,2

reduces cutaneous scarring in adult rodents. Journal of cell science 1994; 107 ( Pt 5):

1137-57.

118 Kohama K, Nonaka K, Hosokawa R et al. TGF-beta-3 promotes scarless repair of

cleft lip in mouse fetuses. Journal of dental research 2002; 81: 688-94.

119 Blanchette F, Day R, Dong W et al. TGFbeta1 regulates gene expression of its own

converting enzyme furin. The Journal of clinical investigation 1997; 99: 1974-83.

120 Yue J, Mulder KM. Transforming growth factor-beta signal transduction in

epithelial cells. Pharmacology & therapeutics 2001; 91: 1-34.

261

121 Crowe MJ, Doetschman T, Greenhalgh DG. Delayed wound healing in

immunodeficient TGF-beta 1 knockout mice. The Journal of investigative

dermatology 2000; 115: 3-11.

122 Saharinen J, Hyytiainen M, Taipale J et al. Latent transforming growth factor-beta

binding proteins (LTBPs)--structural extracellular matrix proteins for targeting

TGF-beta action. Cytokine & growth factor reviews 1999; 10: 99-117.

123 Barcellos-Hoff MH, Dix TA. Redox-mediated activation of latent transforming

growth factor-beta 1. Molecular endocrinology (Baltimore, Md 1996; 10: 1077-83.

124 Todorovic V, Jurukovski V, Chen Y et al. Latent TGF-beta binding proteins. The

international journal of biochemistry & cell biology 2005; 37: 38-41.

125 Taipale J, Miyazono K, Heldin CH et al. Latent transforming growth factor-beta 1

associates to fibroblast extracellular matrix via latent TGF-beta binding protein.

The Journal of cell biology 1994; 124: 171-81.

126 Miyazono K, Olofsson A, Colosetti P et al. A role of the latent TGF-beta 1-binding

protein in the assembly and secretion of TGF-beta 1. The EMBO journal 1991; 10:

1091-101.

127 Miyazono K, Thyberg J, Heldin CH. Retention of the transforming growth factor-

beta 1 precursor in the Golgi complex in a latent endoglycosidase H-sensitive

form. The Journal of biological chemistry 1992; 267: 5668-75.

262

128 Weiskirchen R, Meurer SK, Gressner OA et al. BMP-7 as antagonist of

fibrosis. Front Biosci 2009; 14: 4992-5012.

129 Groppe J, Hinck CS, Samavarchi-Tehrani P et al. Cooperative assembly of TGF-

beta superfamily signaling complexes is mediated by two disparate mechanisms

and distinct modes of receptor binding. Molecular cell 2008; 29: 157-68.

130 Clark DA, Coker R. Transforming growth factor-beta (TGF-beta). The international

journal of biochemistry & cell biology 1998; 30: 293-8.

131 Lopez-Casillas F, Wrana JL, Massague J. Betaglycan presents ligand to the TGF

beta signaling receptor. Cell 1993; 73: 1435-44.

132 Rechtman MM, Nakaryakov A, Shapira KE et al. Different domains regulate

homomeric and heteromeric complex formation among type I and type II

transforming growth factor-beta receptors. The Journal of biological chemistry 2009;

284: 7843-52.

133 Miura Y, Miura O, Ihle JN et al. Activation of the mitogen-activated protein

kinase pathway by the erythropoietin receptor. The Journal of biological chemistry

1994; 269: 29962-9.

134 Derynck R, Feng XH. TGF-beta receptor signaling. Biochimica et biophysica acta

1997; 1333: F105-50.

263

135 Yamashita H, ten Dijke P, Huylebroeck D et al. Osteogenic protein-1 binds to

activin type II receptors and induces certain activin-like effects. The Journal of cell

biology 1995; 130: 217-26.

136 Wrana JL, Attisano L, Wieser R et al. Mechanism of activation of the TGF-beta

receptor. Nature 1994; 370: 341-7.

137 Massague J. TGF-beta signal transduction. Annual review of biochemistry 1998; 67:

753-91.

138 Derynck R, Zhang YE. Smad-dependent and Smad-independent pathways in

TGF-beta family signalling. Nature 2003; 425: 577-84.

139 Piek E, Heldin CH, Ten Dijke P. Specificity, diversity, and regulation in TGF-beta

superfamily signaling. Faseb J 1999; 13: 2105-24.

140 Saha D, Datta PK, Beauchamp RD. Oncogenic ras represses transforming growth

factor-beta /Smad signaling by degrading tumor suppressor Smad4. The Journal of

biological chemistry 2001; 276: 29531-7.

141 Kresse H, Schonherr E. Proteoglycans of the extracellular matrix and growth

control. Journal of cellular physiology 2001; 189: 266-74.

142 Soo C, Hu FY, Zhang X et al. Differential expression of fibromodulin, a

transforming growth factor-beta modulator, in fetal skin development and

scarless repair. The American journal of pathology 2000; 157: 423-33.

264

143 Merline R, Schaefer RM, Schaefer L. The matricellular functions of small leucine-

rich proteoglycans (SLRPs). J Cell Commun Signal 2009.

144 Iozzo RV. The biology of the small leucine-rich proteoglycans. Functional

network of interactive proteins. The Journal of biological chemistry 1999; 274: 18843-

6.

145 Yamaguchi Y, Mann DM, Ruoslahti E. Negative regulation of transforming

growth factor-beta by the proteoglycan decorin. Nature 1990; 346: 281-4.

146 Riquelme C, Larrain J, Schonherr E et al. Antisense inhibition of decorin

expression in myoblasts decreases cell responsiveness to transforming growth

factor beta and accelerates skeletal muscle differentiation. The Journal of biological

chemistry 2001; 276: 3589-96.

147 Nakajima M, Kizawa H, Saitoh M et al. Mechanisms for asporin function and

regulation in articular cartilage. The Journal of biological chemistry 2007; 282: 32185-

92.

148 Hildebrand A, Romaris M, Rasmussen LM et al. Interaction of the small

interstitial proteoglycans biglycan, decorin and fibromodulin with transforming

growth factor beta. Biochem J 1994; 302 ( Pt 2): 527-34.

149 Schonherr E, Broszat M, Brandan E et al. Decorin core protein fragment Leu155-

Val260 interacts with TGF-beta but does not compete for decorin binding to type

I collagen. Arch Biochem Biophys 1998; 355: 241-8.

265

150 Yu Q, Stamenkovic I. Cell surface-localized matrix metalloproteinase-9

proteolytically activates TGF-beta and promotes tumor invasion and

angiogenesis. Genes & development 2000; 14: 163-76.

151 Grainger DJ, Wakefield L, Bethell HW et al. Release and activation of platelet

latent TGF-beta in blood clots during dissolution with plasmin. Nature medicine

1995; 1: 932-7.

152 Bonni A, Brunet A, West AE et al. Cell survival promoted by the Ras-MAPK

signaling pathway by transcription-dependent and -independent mechanisms.

Science (New York, N.Y 1999; 286: 1358-62.

153 Zhang W, Liu HT. MAPK signal pathways in the regulation of cell proliferation

in mammalian cells. Cell Res 2002; 12: 9-18.

154 Huang C, Jacobson K, Schaller MD. MAP kinases and cell migration. Journal of

cell science 2004; 117: 4619-28.

155 Widmann C, Gibson S, Jarpe MB et al. Mitogen-activated protein kinase:

conservation of a three-kinase module from yeast to human. Physiological reviews

1999; 79: 143-80.

156 Brown MD, Sacks DB. Protein scaffolds in MAP kinase signalling. Cellular

signalling 2009; 21: 462-9.

157 Shaw AS, Filbert EL. Scaffold proteins and immune-cell signalling. Nat Rev

Immunol 2009; 9: 47-56.

266

158 Chang L, Karin M. Mammalian MAP kinase signalling cascades. Nature 2001;

410: 37-40.

159 Pearson G, Robinson F, Beers Gibson T et al. Mitogen-activated protein (MAP)

kinase pathways: regulation and physiological functions. Endocrine reviews 2001;

22: 153-83.

160 Vetter IR, Wittinghofer A. The guanine nucleotide-binding switch in three

dimensions. Science (New York, N.Y 2001; 294: 1299-304.

161 Roux PP, Blenis J. ERK and p38 MAPK-activated protein kinases: a family of

protein kinases with diverse biological functions. Microbiol Mol Biol Rev 2004; 68:

320-44.

162 Ichijo H. From receptors to stress-activated MAP kinases. Oncogene 1999; 18:

6087-93.

163 Liu P, Cheng H, Roberts TM et al. Targeting the phosphoinositide 3-kinase

pathway in cancer. Nat Rev Drug Discov 2009; 8: 627-44.

164 Chen RH, Chang MC, Su YH et al. Interleukin-6 inhibits transforming growth

factor-beta-induced apoptosis through the phosphatidylinositol 3-kinase/Akt and

signal transducers and activators of transcription 3 pathways. The Journal of

biological chemistry 1999; 274: 23013-9.

267

165 Fukuda M, Longnecker R. Latent 2A inhibits transforming

growth factor-beta 1-induced apoptosis through the phosphatidylinositol 3-

kinase/Akt pathway. J Virol 2004; 78: 1697-705.

166 Tian B, Lessan K, Kahm J et al. beta 1 integrin regulates fibroblast viability during

collagen matrix contraction through a phosphatidylinositol 3-kinase/Akt/protein

kinase B signaling pathway. The Journal of biological chemistry 2002; 277: 24667-75.

167 Guo X, Wang XF. Signaling cross-talk between TGF-beta/BMP and other

pathways. Cell Res 2009; 19: 71-88.

168 Cordenonsi M, Montagner M, Adorno M et al. Integration of TGF-beta and

Ras/MAPK signaling through p53 phosphorylation. Science (New York, N.Y 2007;

315: 840-3.

169 Liang M, Melchior F, Feng XH et al. Regulation of Smad4 sumoylation and

transforming growth factor-beta signaling by protein inhibitor of activated

STAT1. The Journal of biological chemistry 2004; 279: 22857-65.

170 Brodin G, Ahgren A, ten Dijke P et al. Efficient TGF-beta induction of the Smad7

gene requires cooperation between AP-1, Sp1, and Smad proteins on the mouse

Smad7 promoter. The Journal of biological chemistry 2000; 275: 29023-30.

171 Uchida K, Suzuki H, Ohashi T et al. Involvement of MAP kinase cascades in

Smad7 transcriptional regulation. Biochemical and biophysical research

communications 2001; 289: 376-81.

268

172 Asano Y, Ihn H, Yamane K et al. Phosphatidylinositol 3-kinase is involved in

alpha2(I) collagen gene expression in normal and scleroderma fibroblasts. J

Immunol 2004; 172: 7123-35.

173 Song K, Wang H, Krebs TL et al. Novel roles of Akt and mTOR in suppressing

TGF-beta/ALK5-mediated Smad3 activation. The EMBO journal 2006; 25: 58-69.

174 Wang SE, Shin I, Wu FY et al. HER2/Neu (ErbB2) signaling to Rac1-Pak1 is

temporally and spatially modulated by transforming growth factor beta. Cancer

research 2006; 66: 9591-600.

175 Heinlein CA, Chang C. Androgen receptor (AR) coregulators: an overview.

Endocrine reviews 2002; 23: 175-200.

176 Sampson ER, Yeh SY, Chang HC et al. Identification and characterization of

androgen receptor associated coregulators in prostate cancer cells. Journal of

biological regulators and homeostatic agents 2001; 15: 123-9.

177 Ting HJ, Yeh S, Nishimura K et al. Supervillin associates with androgen receptor

and modulates its transcriptional activity. Proceedings of the National Academy of

Sciences of the United States of America 2002; 99: 661-6.

178 van de Wijngaart DJ, van Royen ME, Hersmus R et al. Novel FXXFF and FXXMF

motifs in androgen receptor cofactors mediate high affinity and specific

interactions with the ligand-binding domain. The Journal of biological chemistry

2006; 281: 19407-16.

269

179 Safiejko-Mroczka B, Bell PB, Jr. Reorganization of the actin cytoskeleton in the

protruding lamellae of human fibroblasts. Cell motility and the cytoskeleton 2001;

50: 13-32.

180 Kopecki Z, Cowin AJ. Flightless I: an actin-remodelling protein and an important

negative regulator of wound repair. The international journal of biochemistry & cell

biology 2008; 40: 1415-9.

181 Chen W, Wahl SM. Manipulation of TGF-beta to control autoimmune and

chronic inflammatory diseases. Microbes Infect 1999; 1: 1367-80.

182 Bandyopadhyay B, Fan J, Guan S et al. A "traffic control" role for TGFbeta3:

orchestrating dermal and epidermal cell motility during wound healing. The

Journal of cell biology 2006; 172: 1093-105.

183 Colwell AS, Faudoa R, Krummel TM et al. Transforming growth factor-beta,

Smad, and collagen expression patterns in fetal and adult keratinocytes. Plastic

and reconstructive surgery 2007; 119: 852-7.

184 Colwell AS, Krummel TM, Longaker MT et al. Fetal and adult fibroblasts have

similar TGF-beta-mediated, Smad-dependent signaling pathways. Plastic and

reconstructive surgery 2006; 117: 2277-83.

185 Gailit J, Welch MP, Clark RA. TGF-beta 1 stimulates expression of keratinocyte

integrins during re-epithelialization of cutaneous wounds. The Journal of

investigative dermatology 1994; 103: 221-7.

270

186 Zambruno G, Marchisio PC, Marconi A et al. Transforming growth factor-beta 1

modulates beta 1 and beta 5 integrin receptors and induces the de novo

expression of the alpha v beta 6 heterodimer in normal human keratinocytes:

implications for wound healing. The Journal of cell biology 1995; 129: 853-65.

187 Sellheyer K, Bickenbach JR, Rothnagel JA et al. Inhibition of skin development by

overexpression of transforming growth factor beta 1 in the epidermis of

transgenic mice. Proceedings of the National Academy of Sciences of the United States

of America 1993; 90: 5237-41.

188 Adams DH, Strudwick XL, Kopecki Z et al. Gender specific effects on the actin-

remodelling protein Flightless I and TGF-beta1 contribute to impaired wound

healing in aged skin. The international journal of biochemistry & cell biology 2008; 40:

1555-69.

189 Arkell RM, Cadman M, Marsland T et al. Genetic, physical, and phenotypic

characterization of the Del(13)Svea36H mouse. Mamm Genome 2001; 12: 687-94.

190 Brown RL, Ormsby I, Doetschman TC et al. Wound healing in the transforming

growth factor-beta-deficient mouse. Wound Repair Regen 1995; 3: 25-36.

191 Letterio JJ, Roberts AB. Transforming growth factor-beta1-deficient mice:

identification of isoform-specific activities in vivo. J Leukoc Biol 1996; 59: 769-74.

192 Shah M, Foreman DM, Ferguson MW. Control of scarring in adult wounds by

neutralising antibody to transforming growth factor beta. Lancet 1992; 339: 213-4.

271

193 Koch RM, Roche NS, Parks WT et al. Incisional wound healing in transforming

growth factor-beta1 null mice. Wound Repair Regen 2000; 8: 179-91.

194 Spindler KP, Murray MM, Detwiler KB et al. The biomechanical response to

doses of TGF-beta 2 in the healing rabbit medial collateral ligament. J Orthop Res

2003; 21: 245-9.

195 Lu L, Saulis AS, Liu WR et al. The temporal effects of anti-TGF-beta1, 2, and 3

monoclonal antibody on wound healing and hypertrophic scar formation. Journal

of the American College of Surgeons 2005; 201: 391-7.

196 Hosokawa R, Nonaka K, Morifuji M et al. TGF-beta 3 decreases type I collagen

and scarring after labioplasty. Journal of dental research 2003; 82: 558-64.

197 Buchanan SG, Gay NJ. Structural and functional diversity in the leucine-rich

repeat family of proteins. Progress in biophysics and molecular biology 1996; 65: 1-44.

198 Wang T, Chuang TH, Ronni T et al. Flightless I homolog negatively modulates

the TLR pathway. J Immunol 2006; 176: 1355-62.

199 Claudianos C, Campbell HD. The novel flightless-I gene brings together two

gene families, actin-binding proteins related to gelsolin and leucine-rich-repeat

proteins involved in Ras signal transduction. Mol Biol Evol 1995; 12: 405-14.

200 Wrana JL, Attisano L, Carcamo J et al. TGF beta signals through a heteromeric

protein kinase receptor complex. Cell 1992; 71: 1003-14.

201 Moustakas A. Smad signalling network. Journal of cell science 2002; 115: 3355-6.

272

202 Shi Y, Massague J. Mechanisms of TGF-beta signaling from cell membrane to the

nucleus. Cell 2003; 113: 685-700.

203 Kim SJ, Angel P, Lafyatis R et al. Autoinduction of transforming growth factor

beta 1 is mediated by the AP-1 complex. Molecular and cellular biology 1990; 10:

1492-7.

204 Wakefield LM, Piek E, Bottinger EP. TGF-beta signaling in mammary gland

development and tumorigenesis. J Mammary Gland Biol Neoplasia 2001; 6: 67-82.

205 Sano Y, Tokitou F, Dai P et al. CBP alleviates the intramolecular inhibition of

ATF-2 function. The Journal of biological chemistry 1998; 273: 29098-105.

206 Beauchamp RD, Sheng HM, Ishizuka J et al. Transforming growth factor (TGF)-

beta stimulates hepatic jun-B and fos-B proto-oncogenes and decreases albumin

mRNA. Annals of surgery 1992; 216: 300-7; discussion 7-8.

207 Thorp BH, Anderson I, Jakowlew SB. Transforming growth factor-beta 1, -beta 2

and -beta 3 in cartilage and bone cells during endochondral ossification in the

chick. Development 1992; 114: 907-11.

208 Zhang S, Fei T, Zhang L et al. Smad7 antagonizes transforming growth factor

beta signaling in the nucleus by interfering with functional Smad-DNA complex

formation. Molecular and cellular biology 2007; 27: 4488-99.

273

209 Deliot N, Chavent M, Nourry C et al. Biochemical studies and molecular

dynamics simulations of Smad3-Erbin interaction identify a non-classical Erbin

PDZ binding. Biochemical and biophysical research communications 2009; 378: 360-5.

210 Kolch W. Erbin: sorting out ErbB2 receptors or giving Ras a break? Sci STKE

2003; 2003: pe37.

211 Dai F, Chang C, Lin X et al. Erbin inhibits transforming growth factor beta

signaling through a novel Smad-interacting domain. Molecular and cellular biology

2007; 27: 6183-94.

212 Itoh S, Landstrom M, Hermansson A et al. Transforming growth factor beta1

induces nuclear export of inhibitory Smad7. The Journal of biological chemistry

1998; 273: 29195-201.

213 Dore JJ, DeWitt JC, Setty N et al. Multiple signaling pathways converge to

regulate bone-morphogenetic-protein-dependent glial gene expression. Dev

Neurosci 2009; 31: 473-86.

214 Cakir M, Grossman AB. Targeting MAPK (Ras/ERK) and PI3K/Akt pathways in

pituitary tumorigenesis. Expert Opin Ther Targets 2009; 13: 1121-34.

215 Conery AR, Cao Y, Thompson EA et al. Akt interacts directly with Smad3 to

regulate the sensitivity to TGF-beta induced apoptosis. Nature cell biology 2004; 6:

366-72.

274

216 Remy I, Montmarquette A, Michnick SW. PKB/Akt modulates TGF-beta

signalling through a direct interaction with Smad3. Nature cell biology 2004; 6:

358-65.

217 Faissner A, Heck N, Dobbertin A et al. DSD-1-Proteoglycan/Phosphacan and

receptor protein tyrosine phosphatase-beta isoforms during development and

regeneration of neural tissues. Advances in experimental medicine and biology 2006;

557: 25-53.

218 Ramaswamy S, Nakamura N, Vazquez F et al. Regulation of G1 progression by

the PTEN tumor suppressor protein is linked to inhibition of the

phosphatidylinositol 3-kinase/Akt pathway. Proceedings of the National Academy of

Sciences of the United States of America 1999; 96: 2110-5.

219 Kandel ES, Skeen J, Majewski N et al. Activation of Akt/protein kinase B

overcomes a G(2)/m cell cycle checkpoint induced by DNA damage. Molecular

and cellular biology 2002; 22: 7831-41.

220 Masuelli L, Cutler ML. Increased expression of the Ras suppressor Rsu-1

enhances Erk-2 activation and inhibits Jun kinase activation. Molecular and cellular

biology 1996; 16: 5466-76.

221 Hartsough MT, Mulder KM. Transforming growth factor-beta signaling in

epithelial cells. Pharmacology & therapeutics 1997; 75: 21-41.

275

222 Huang SS, Huang JS. TGF-beta control of cell proliferation. J Cell Biochem 2005;

96: 447-62.

223 Acharya PS, Majumdar S, Jacob M et al. Fibroblast migration is mediated by

CD44-dependent TGF beta activation. Journal of cell science 2008; 121: 1393-402.

224 Hayashi H, Abdollah S, Qiu Y et al. The MAD-related protein Smad7 associates

with the TGFbeta receptor and functions as an antagonist of TGFbeta signaling.

Cell 1997; 89: 1165-73.

225 Nakao A, Afrakhte M, Moren A et al. Identification of Smad7, a TGFbeta-

inducible antagonist of TGF-beta signalling. Nature 1997; 389: 631-5.

226 Desmouliere A, Geinoz A, Gabbiani F et al. Transforming growth factor-beta 1

induces alpha-smooth muscle actin expression in granulation tissue

myofibroblasts and in quiescent and growing cultured fibroblasts. The Journal of

cell biology 1993; 122: 103-11.

227 Lijnen P, Petrov V, Rumilla K et al. Stimulation of collagen gel contraction by

angiotensin II and III in cardiac fibroblasts. J Renin Angiotensin Aldosterone Syst

2002; 3: 160-6.

228 Diegelmann RF, Evans MC. Wound healing: an overview of acute, fibrotic and

delayed healing. Front Biosci 2004; 9: 283-9.

229 Bain J, Plater L, Elliott M et al. The selectivity of protein kinase inhibitors: a

further update. Biochem J 2007; 408: 297-315.

276

230 Javelaud D, Mauviel A. Crosstalk mechanisms between the mitogen-activated

protein kinase pathways and Smad signaling downstream of TGF-beta:

implications for carcinogenesis. Oncogene 2005; 24: 5742-50.

231 Kottler UB, Junemann AG, Aigner T et al. Comparative effects of TGF-beta 1 and

TGF-beta 2 on extracellular matrix production, proliferation, migration, and

collagen contraction of human Tenon's capsule fibroblasts in pseudoexfoliation

and primary open-angle glaucoma. Exp Eye Res 2005; 80: 121-34.

232 Ellis I, Grey AM, Schor AM et al. Antagonistic effects of TGF-beta 1 and MSF on

fibroblast migration and hyaluronic acid synthesis. Possible implications for

dermal wound healing. Journal of cell science 1992; 102 ( Pt 3): 447-56.

233 Junttila MR, Li SP, Westermarck J. Phosphatase-mediated crosstalk between

MAPK signaling pathways in the regulation of cell survival. Faseb J 2008; 22: 954-

65.

234 Uttamsingh S, Bao X, Nguyen KT et al. Synergistic effect between EGF and TGF-

beta1 in inducing oncogenic properties of intestinal epithelial cells. Oncogene

2008; 27: 2626-34.

235 Han MY, Kosako H, Watanabe T et al. Extracellular signal-regulated

kinase/mitogen-activated protein kinase regulates actin organization and cell

motility by phosphorylating the actin cross-linking protein EPLIN. Molecular and

cellular biology 2007; 27: 8190-204.

277

236 Urbanucci A, Waltering KK, Suikki HE et al. Androgen regulation of the

androgen receptor coregulators. BMC cancer 2008; 8: 219.

237 Lee SB, Xuan Nguyen TL, Choi JW et al. Nuclear Akt interacts with B23/NPM

and protects it from proteolytic cleavage, enhancing cell survival. Proceedings of

the National Academy of Sciences of the United States of America 2008; 105: 16584-9.

278