investigating the effects of, the isoflavone, phenoxodiol

281
INVESTIGATING THE CYTOTOXIC EFFECTS OF THE ISOFLAVONE, PHENOXODIOL, ON PROSTATE CANCER CELLS. This thesis is presented for the degree of Doctor of Philosophy The University of Western Australia School of Anatomy, Physiology and Human Biology Simon David Mahoney, BSc Hons 2012 Supervised by: Professor Arunasalam Dharmarajan Professor Michael Millward

Upload: others

Post on 23-Feb-2022

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Investigating the effects of, the isoflavone, Phenoxodiol

i

INVESTIGATING THE CYTOTOXIC

EFFECTS OF THE ISOFLAVONE,

PHENOXODIOL, ON PROSTATE

CANCER CELLS.

This thesis is presented for the degree of

Doctor of Philosophy

The University of Western Australia

School of Anatomy, Physiology and Human Biology

Simon David Mahoney, BSc Hons

2012

Supervised by:

Professor Arunasalam Dharmarajan

Professor Michael Millward

Page 2: Investigating the effects of, the isoflavone, Phenoxodiol

ii

THIS THESIS IS DEDICATED TO MY FAMILY, MY FRIENDS AND MY SUPERVISORS, WHO’S

SUPPORT WAS UNWAVERING.

Page 3: Investigating the effects of, the isoflavone, Phenoxodiol

iii

DECLARATION

The work presented within this thesis was completed between March 2005 and June 2012 in

the School of Anatomy, Physiology and Human Biology at the University of Western Australia. I

hereby declare that all work presented is entirely my own, unless explicitly stated otherwise.

All contributions by others are formally disclosed and duly acknowledged.

Simon Mahoney

1st July 2012

Page 4: Investigating the effects of, the isoflavone, Phenoxodiol

iv

ACKNOWLEGEMENTS

I would like to acknowledge the support of the following people without whom my thesis

would not have been possible.

My supervisors Professor Arun Dharmarajan and Professory Michael Millward

Mr Greg Cozens, Dr Peter Mark, Dr Kathy Heel and Ms Susan Hisheh for their technical

expertise and support.

My fellow students; Clare Berry, Lloyd White, Jeremy Drake, Bijanka Franklin, Brilliana Von

Katterfeld, Mats Hellstrom, Jill Muhling, Hannah Crabb, Margaret Pollet, Melissa Berg, Thea

Shavlakadze, Kasie Jeffrey for their expertise and support.

My Family, Friends and anyone I may have missed for all their support and help throughout the

process and Kerry for helping me through the last obstacles.

Page 5: Investigating the effects of, the isoflavone, Phenoxodiol

v

CONTENTS

Declaration .................................................................................................................... iii

Acknowlegements ....................................................................................................... iv

List of Equations ............................................................................................................ xi

List of Figures ................................................................................................................ xi

List of Tables ............................................................................................................... xiv

List of Abbreviations .................................................................................................... xv

Thesis Layout ............................................................................................................... xix

Abstract ....................................................................................................................... xix

1. Chapter One: Introduction ................................................................................. 1

2. Chapter Two: Literature Review ......................................................................... 4

2.1. Cancer ........................................................................................................................... 4

2.1.1. Common Cancers and Prostate Cancer In Males in Australia ............................... 4

2.1.2. Cancer As a Disease In Australia ........................................................................... 6

2.2. Phenoxodiol .................................................................................................................. 7

2.2.1. Flavanoids and Isoflavones ................................................................................... 8

2.2.2. Phenoxodiols Reported Method Of Action ......................................................... 10

2.2.3. Synergistic Activty ............................................................................................... 13

2.2.4. Treatment With Phenoxodiol.............................................................................. 13

2.3. Carcinogenesis / Cancer Development ....................................................................... 14

2.3.1. Evasion Of Apoptosis .......................................................................................... 17

2.3.2. Self-Sufficiency In Growth Signals ....................................................................... 17

2.3.3. Insensitivity To Anti-Growth Signals ................................................................... 18

2.3.4. Limitless Replicative Potential ............................................................................ 19

2.3.5. Sustained Angiogenesis ....................................................................................... 20

Page 6: Investigating the effects of, the isoflavone, Phenoxodiol

vi

2.3.6. Tissue Invasion And Metastasis ........................................................................... 20

2.3.7. Pluripotent Stem Cell Differentiation and Development .................................... 21

2.3.8. Gap Junctional Intercellular Connections ............................................................ 22

2.4. The Prostate ................................................................................................................ 23

2.4.1. Testing For Prostate Cancer ................................................................................ 25

2.4.2. Phenotypic Progression of Prostate Cancer ........................................................ 25

2.4.3. Treatment of Prostate Cancer ............................................................................. 27

2.4.4. Chemotherapy and Prostate Cancer ................................................................... 28

2.5. Apoptosis ..................................................................................................................... 29

2.5.1. Morphology of Apoptotic Cells ............................................................................ 31

2.5.2. Apoptotic Signalling ............................................................................................. 33

2.5.3. Extrinsic Pathways ............................................................................................... 34

2.5.4. Intrinsic Pathways ................................................................................................ 35

2.5.5. Caspases .............................................................................................................. 36

2.5.6. Apoptosome ........................................................................................................ 38

2.5.7. Bcl-2 Family .......................................................................................................... 38

2.5.8. X-Linked Inhibitor of Apoptosis Protein .............................................................. 39

2.5.9. Apoptosis Inducing Factor ................................................................................... 40

2.5.10. Reactive Oxygen Species ..................................................................................... 41

2.6. The Cell Cycle ............................................................................................................... 42

2.6.1. Cyclins, Cyclin Dependant Kinases ....................................................................... 43

2.6.2. p21WAF1/p53 ......................................................................................................... 43

2.6.3. Ki-67 ..................................................................................................................... 44

Page 7: Investigating the effects of, the isoflavone, Phenoxodiol

vii

2.6.4. c-Myc ................................................................................................................... 45

2.7. Wnt Signalling ............................................................................................................. 45

2.7.1. β-catenin ............................................................................................................. 46

2.7.2. Androgen Receptor and β-catenin ...................................................................... 47

2.7.3. sFRP4 ................................................................................................................... 48

3. Chapter Three: Materials and Methods ............................................................ 50

3.1. Tissue Culture .............................................................................................................. 50

3.2. Cell culture .................................................................................................................. 50

3.2.1. Subculture And Counting Of Cells ....................................................................... 51

3.2.2. Cell Cryopreservation and Storage ..................................................................... 51

3.2.3. LNCaP Cells .......................................................................................................... 52

3.2.3.1 LNCaP media formulation ............................................................................... 53

3.2.4. DU145 Cells ......................................................................................................... 53

3.2.4.1 DU145 Media Formulation ............................................................................. 53

3.2.5. PC3 Cells .............................................................................................................. 54

3.2.5.1 PC3 Media Formulation .................................................................................. 54

3.3. Drug Dilutions ............................................................................................................. 54

3.3.1. Phenoxodiol ........................................................................................................ 54

3.3.2. Phenoxodiol Working Stock ................................................................................ 55

3.3.3. Phenoxodiol Working Solution ........................................................................... 55

3.3.4. Phenoxodiol Treatment ...................................................................................... 57

3.3.5. Docetaxel ............................................................................................................ 57

3.3.6. Docetaxel Working Stock and Working Solution ................................................ 57

3.4. Optimisation of Proliferation ...................................................................................... 58

Page 8: Investigating the effects of, the isoflavone, Phenoxodiol

viii

3.4.1. Determination of Cell Seeding Concentrations ................................................... 58

3.4.2. Phenoxodiol MTS Assay ....................................................................................... 59

3.4.3. Phenoxodiol Docetaxel Isobologram ................................................................... 61

3.4.4. Phenoxodiol And Caspase Inhibition ................................................................... 64

3.4.5. Phenoxodiol and Docetaxel Combination Treatment ......................................... 65

3.4.6. Phenoxodiol and Purified sFRP4 Protein Combination Treatment ..................... 66

3.5. Reactive Oxygen Species Detection............................................................................. 67

3.6. Acidity Analysis ............................................................................................................ 69

3.7. Apoptosis Analysis Assays ........................................................................................... 69

3.7.1. DNA Extraction .................................................................................................... 69

3.7.2. 3’-End Labelling DNA Fragmentation Analysis .................................................... 71

3.7.3. Annexin-V-Fluos Propidium Iodide Flow Cyometry ............................................. 72

3.7.4. Sybr Gold Fragmentation Analysis ....................................................................... 75

3.7.5. JC-1 Mitochondrial Potential Assay ..................................................................... 76

3.7.6. Caspase-3 Activity Assay ...................................................................................... 78

3.8. Cell Cycle Analysis ........................................................................................................ 79

3.8.1. Cell Cycle Preparation .......................................................................................... 80

3.8.2. Cell Cycle Flow Cytometry ................................................................................... 81

3.8.3. Cell Cycle Data Analysis ....................................................................................... 85

3.9. Assessment of RNA Expression ................................................................................... 87

3.9.1. RNA Extraction ..................................................................................................... 88

3.9.2. RNA Integrity ....................................................................................................... 89

3.9.3. DNAse Treatment ................................................................................................ 89

Page 9: Investigating the effects of, the isoflavone, Phenoxodiol

ix

3.9.4. Reverse Transcriptase Polymerase Chain Reaction ............................................ 90

3.9.5. Post PCR Clean-Up .............................................................................................. 91

3.9.6. Primer Design ...................................................................................................... 92

3.9.7. Real Time Quantitative PCR Analysis .................................................................. 93

3.9.8. Gel Electrophoresis and Extraction ..................................................................... 96

3.9.9. qPCR Standard Production .................................................................................. 97

3.10. Assessment of Protein Expression ........................................................................ 100

3.10.1. Protein Extraction ............................................................................................. 101

3.10.2. Bradford Protein Quantification Assay ............................................................. 102

3.10.3. SDS-PAGE Western Blot Analysis ...................................................................... 103

3.10.4. Immunoblotting ................................................................................................ 107

3.10.5. Quantification of Western Blot Analysis ........................................................... 109

3.11. Statistical Analysis ................................................................................................. 110

4. Chapter Four: The Cytotoxic Effects of Phenoxodiol On The Prostate Cancer Cell

Lines; LNCaP, DU145 and PC3 ................................................................................ 111

4.1. Introduction .............................................................................................................. 111

4.2. Aims........................................................................................................................... 114

4.3. Methodology ............................................................................................................. 114

4.4. Results ....................................................................................................................... 118

4.5. Discussion .................................................................................................................. 134

5. Chapter Five: Cell Death Signalling In Response To Phenoxodiol Treatment .... 138

5.1. Introduction .............................................................................................................. 138

5.2. Aims........................................................................................................................... 142

Page 10: Investigating the effects of, the isoflavone, Phenoxodiol

x

5.3. Methodology ............................................................................................................. 142

5.4. Results ....................................................................................................................... 145

5.5. Discussion .................................................................................................................. 166

6. Chapter Six: The Effects of Phenoxodiol On The Cell Cycle .............................. 172

6.1. Introduction ............................................................................................................... 172

6.2. Aims ........................................................................................................................... 177

6.3. Methodology ............................................................................................................. 177

6.4. Results ....................................................................................................................... 179

6.5. Discussion .................................................................................................................. 197

7. Chapter Seven: Phenoxodiol in Combination with Docetaxel ......................... 203

7.1. Introduction ............................................................................................................... 203

7.2. Aims ........................................................................................................................... 207

7.3. Methodology ............................................................................................................. 207

7.4. Results ....................................................................................................................... 210

7.5. Discussion .................................................................................................................. 225

8. Chapter Eight: General Discussion.................................................................. 230

8.1. Discussion .................................................................................................................. 230

8.2. Conclusion ................................................................................................................. 239

8.3. Limitations ................................................................................................................. 242

9. Bibliography .................................................................................................. 244

10. Appendices ................................................................................................ 261

Page 11: Investigating the effects of, the isoflavone, Phenoxodiol

xi

LIST OF EQUATIONS

Equation 1 Phenoxodiol Working Stock Equation ...................................................................... 55

Equation 2 Phenoxodiol Working Solution Equation ................................................................. 56

Equation 3 Docetaxel Stock And Working Solutions .................................................................. 58

LIST OF FIGURES

Figure 1: Burden of Disease in Australia In 2003 .......................................................................... 6

Figure 2 Phenoxodiol Molecular Structure ................................................................................... 8

Figure 3 Production of Equol From Daidzein ................................................................................ 9

Figure 4 Phenoxodiols Reported Method of Action As Adapted From Silasi et al. 2009 ........... 12

Figure 5: Acquired Capabilities of Cancer Adapted from Hanahan et al 2000 and Trosko et al

2004 ............................................................................................................................................ 16

Figure 6: Stem Cell Theory Adapted from Trosko et al 2004 ...................................................... 22

Figure 7 Characteristic Progression of Prostate Cancer From Early to Late Stage ..................... 26

Figure 8 Development Of Resistance To Chemotherapeutic Treatment Adapted from

Johnstone, Ruefli and Lowe 2002 ............................................................................................... 29

Figure 9 Simple Apoptotic Signalling Adapted From Riedl and Salvesen 2007 .......................... 37

Figure 10 MTS Assay Setup ......................................................................................................... 60

Figure 11 96 Well isobologram Setup ......................................................................................... 62

Figure 12 LNCaP Sybr Gold Visualised DNA Fragmentation With DNA Ladder, Low Weight DNA

Fragmentation Is Visible ............................................................................................................. 76

Figure 13 Cell Populations Multi-Gated In FACS Diva Software For acquisition And Analysis ... 83

Figure 14 Cell Cycle Populations For Analysis Aquired Using Gated Populations In FACS Diva

Software ...................................................................................................................................... 84

Figure 15 Example Of Cell Population Gating For Analysis In FlowJo Software ......................... 85

Page 12: Investigating the effects of, the isoflavone, Phenoxodiol

xii

Figure 16 Watson versus Dean-Jett-Fox Cell Cycle Analysis In FlowJo Software ........................ 86

Figure 17 Watson Pragmatic Cell Cycle Analysis of PC3 Cells Undergoing Phenoxodiol

Treatment .................................................................................................................................... 87

Figure 18 Example Of A Cyclin-D1 Melt Curve ............................................................................ 96

Figure 19 qPCR Standards Cycling Assessment ........................................................................... 98

Figure 20 Standards Analysis by Corbett Rotor-Gene 6000 Software. ....................................... 99

Figure 21 qPCR Full Run Standards and Samples ........................................................................ 99

Figure 22 Standards and Samples Analysis By Corbett Rotor-Gene 6000 Software ................. 100

Figure 23 Cell Proliferation Rates At 6, 24 and 48 hours .......................................................... 119

Figure 24 Cell Proliferation Rates After 24 and 48 Hours Of Phenoxodiol Treatment ............. 121

Figure 25 pH Changes In Phenoxodiol Treated Culture Media ................................................. 122

Figure 26 JC-1 Analysis of Mitochondrial Membrane Potential Over 24 and 48 Hours of

Treatment .................................................................................................................................. 125

Figure 27 Fluorescent Analysis of Caspase-3 Activity after Phenoxodiol Treatment Over 24 and

48 Hours .................................................................................................................................... 127

Figure 28 3'-End Labelling Apoptotic Analysis Post Phenoxodiol Treatment over 24 and 48

Hours ......................................................................................................................................... 129

Figure 29 An Example of LNCaP 3’-end Labelling Qualitative DNA Laddering After Exposure to

Film ............................................................................................................................................ 130

Figure 30 Annexin V-Fluos / Propidium Iodide Double Staining Analysis of Prostate Cancer Cells

Post Phenoxodiol Treatment over 24 and 48 Hours ................................................................. 133

Figure 31 AIF mRNA Expression Analysis of Prostate Cancer Cells Post Phenoxodiol Treatment

Over 24 And 48 Hours ............................................................................................................... 146

Figure 32 Bax mRNA Expression Analysis of Prostate Cancer Cells Post Phenoxodiol Treatment

Over 24 And 48 Hours ............................................................................................................... 148

Figure 33 Bcl-xL mRNA Expression Analysis of Prostate Cancer Cells Post Phenoxodiol

Treatment Over 24 And 48 Hours ............................................................................................. 150

Page 13: Investigating the effects of, the isoflavone, Phenoxodiol

xiii

Figure 34 Caspase-3 mRNA Expression Analysis of Prostate Cancer Cells Post Phenoxodiol

Treatment Over 24 And 48 Hours ............................................................................................. 152

Figure 35 xIAP mRNA Expression Analysis of Prostate Cancer Cells Post Phenoxodiol Treatment

Over 24 And 48 Hours ............................................................................................................... 154

Figure 36 AIF Protein Level Analysis Of Prostate Cancer Cells Post Phenoxodiol Treatment Over

24 And 48 Hours........................................................................................................................ 156

Figure 37 Bax Protein Level Analysis Of Prostate Cancer Cells Post Phenoxodiol Treatment Over

24 And 48 Hours........................................................................................................................ 157

Figure 38 Bcl-xL Protein Level Analysis Of Prostate Cancer Cells Post Phenoxodiol Treatment

Over 24 And 48 Hours ............................................................................................................... 159

Figure 39 xIAP Protein Level Analysis Of Prostate Cancer Cells Over 24 and 48 Hours Post

Phenoxodiol Treatment Over 24 And 48 Hours ........................................................................ 161

Figure 40 Caspase Inhibition Treatment With 10µM Z-VAD-FMK (CI) and 30µM Phenoxodiol

(PXD) Over 48 hours .................................................................................................................. 163

Figure 41 Nitric Oxide Formation In Prostate Cancer Cells Over 24 and 48 Hours Post

Phenoxodiol Treatment Measured Via Griess Assay ................................................................ 165

Figure 42 LNCaP Cell Cycle Analysis After 24 and 48 Hours Of 10µM and 30µM Phenoxodiol

Treatment ................................................................................................................................. 180

Figure 43 DU145 Cycle Analysis After 24 and 48 Hours Of 10µM and 30µM Phenoxodiol

Treatment ................................................................................................................................. 182

Figure 44 PC3 Cell Cycle Analysis After 24 and 48 Hours Of 10µM and 30µM Phenoxodiol

Treatment. ................................................................................................................................ 184

Figure 45 c-Myc mRNA Expression Analysis of Prostate Cancer Over 24 and 48 Hours Post

Phenoxodiol Treatment. ........................................................................................................... 186

Figure 46 Cyclin-D1 mRNA Expression Analysis of Prostate Cancer Cells Over 24 and 48 Hours

Post Phenoxodiol Treatment. ................................................................................................... 188

Page 14: Investigating the effects of, the isoflavone, Phenoxodiol

xiv

Figure 47 Ki-67 mRNA Expression Analysis of Prostate Cancer Cells Over 24 and 48 Hours Post

Phenoxodiol Treatment. ............................................................................................................ 190

Figure 48 p21 mRNA Expression Analysis of Prostate Cancer Over 24 and 48 Hours Post

Phenoxodiol Treatment. ............................................................................................................ 192

Figure 49 Active β-catenin Protein Level Analysis of Prostate Cancer Cells Over 24 and 48 Hours

Post Phenoxodiol Treatment. .................................................................................................... 194

Figure 50 sFRP4 Protein Level Analysis of Prostate Cancer Cells Over 24 and 48 Hours Post

Phenoxodiol Treatment ............................................................................................................. 196

Figure 51 Docetaxel Response Curve Measured After 48 Hours of Treatment. ....................... 212

Figure 52 LNCaP Phenoxodiol / Docetaxel Isobologram Measured After 48 Hours of Treatment

................................................................................................................................................... 213

Figure 53 DU145 Phenoxodiol / Docetaxel Isobologram Measured After 48 Hours Of

Treatment. ................................................................................................................................. 214

Figure 54 PC3 Phenoxodiol / Docetaxel Isobologram Measured After 48 Hours Of Treatment.

................................................................................................................................................... 215

Figure 55 10µM Phenoxodiol 100nM Docetaxel Combination Therapy Measured After 48

Hours Of Treatment. .................................................................................................................. 218

Figure 56 30µM Phenoxodiol 100nM Docetaxel Combination Therapy Measured After 48

Hours Of Treatment. .................................................................................................................. 221

Figure 57 30µM PXD And 500pg/mL sFRP4 Protein Combination Therapy After 48 Hours. .... 224

LIST OF TABLES

Table 1: Predicted Prostate Cancer Rates 2011-2020, Adapted From Begg, 2007 ....................... 5

Table 2 The 10 Most Commonly Occuring Cancer In Australia 2007, Adapted From Begg 2007 . 5

Table 3 Phenoxodiol Treatment Calculations Per mL Of Media ................................................. 56

Table 4: Seeding Concentrations And Volumes .......................................................................... 59

Table 5 Isobologram Setup For Vehicle Control And Treatment Groups .................................... 63

Page 15: Investigating the effects of, the isoflavone, Phenoxodiol

xv

Table 6 Caspase Inhibition and Phenoxodiol Treatment Solutions ............................................ 65

Table 7 sFRP4 & Phenoxodiol Treatment Solutions ................................................................... 67

Table 8 3'-End Labelling Reaction Mixture ................................................................................. 71

Table 9 Cell Staining With Annexin-V/Propidium Iodide (AV/PI) Double Stain .......................... 73

Table 10 Cell Staining Setup For AV/PI ....................................................................................... 74

Table 11 AV/PI Labelling Solutions ............................................................................................. 74

Table 12 Propidium Iodide Staining Solution Per 30mL ............................................................. 81

Table 13 RQ1 DNase Treatment Reaction Components ............................................................. 90

Table 14 RT PCR Reaction Components ..................................................................................... 91

Table 15 Primer Sequences, Product Size And Annealing Temperature .................................... 93

Table 16 Immolase Taq qPCR Reaction Mix ............................................................................... 95

Table 17 qPCR Standards Serial Dilution .................................................................................... 98

Table 18 Bradford Protein Standards Setup ............................................................................. 102

Table 19 SDS-Page/Acrylamide Gel Recipe ............................................................................... 104

Table 20 Example Of Protein Sample Preparation ................................................................... 105

Table 21 Antibody Concentrations ........................................................................................... 108

Table 22 Immunoblotting Protocol ........................................................................................... 109

Table 23 Phenoxodiol/Docetaxel Concentration Combinations For An Isobologram .............. 208

LIST OF ABBREVIATIONS

Abbreviation Extended Form

µ, µm, µM Micro, micrometer, micromolar

AAT Androgen Ablation Therapy

AIF Apoptosis Inducing Factor

APS Ammonium Persulfate

AR Androgen Receptor

ARE Antioxidant Response Element

ATCC American Type Culture Collection

Page 16: Investigating the effects of, the isoflavone, Phenoxodiol

xvi

AV, AVF, Annexin-V Annexin-V-Fluorescein

Bax Bcl-2-associated X Protein

Bcl-2 B-cell Lymphoma 2

Bcl-xL B-cell lymphoma extra large

BLAST Basic Logical Alignment Search Tool

BSA Bovine Serum Albumin

BP Base Pair

BPH Benign Prostatic Hyperplasia

Cat# Catalogue Number

Cdk Cyclin Dependant Kinase

CdkI Cyclin Dependant Kinase Inhibitor

cDNA Complementary DNA

CHAPS 3-[(3-Cholamidopropyl)dimethylammonio]-1-propanesulfonate

CI Caspase Inhibitor (Z-VAD-FMK)

CNS Central Nervous System

CRPC Castrate Resistant Prostate Cancer

DALY Disability Adjusted Life Years

ddATP Dideoxyadenosine Triphosphate

ddH2O Double distilled water / also deionised water

DEAN Diethylamine NONOate diethylammonium salt

DEPC Diethylpyrocarbonate

DHT 5-alpha-dihydrotestosterone

DISC Death-Inducing Signalling Complex

DJF Dean-Jett-Fox

DMDC Dimethyl dicarbonate

DMSO Dimethyl Sulfoxide, (CH3)2SO

DNA Deoxyribonucleic Acid

DOC Docetaxel

DTT Dithiothreitol

EDTA Ethylenediaminetetraacetic Acid

EGTA Ethyleneglyocoltetraacetic Acid

ER Estrogen Receptor

FACS Fluorescence-Activated Cell Sorting

FBS Fetal Bovine Serum

Page 17: Investigating the effects of, the isoflavone, Phenoxodiol

xvii

FCCP Carbonylcyanide-4-trifluoromethoxyphenylhydrazone

FLIP Flice Like Inhibitory Protein

FLUOS Fluorescein

FSC, -A, -H Forward Scatter, -Area, -Height

FZD Frizzled Receptor

g RCF (see below) = 0.00001118rN2 r=rotational radius cm

N=rotating speed in revolutions per minute (RPM)

GIJC Gap Junctional Intercellular Connections

GSK3-β Glycogen Synthase Kinase 3 - Beta

HEPES 4-(2-Hydroxyethyl)piperazine-1-ethanesulfonic acid

HRP Streptavidin-Horse Radish Peroxidase

HRPC Hormone refractory prostate cancer

LEF Lymphoid Enhancer Factor

LRP5/LRP6 Low Density Lipoprotein Receptor-Related Protein

MEM Minimum Essential Medium

M, mM Molar, millimolar

M-MLV Moloney Murine Leukaemia Virus

mRNA Messenger Ribonucleic Acid

MTS 3-(4,5-dimethylthiazol-2-yl)-5-(3-carboxymethoxyphenyl)-

2-(4-sulfophenyl)-2H-tetrazolium

NED N-1-naphthylethylenediamine dichloride

nm, nM Nanometre, nanomolar

NO Nitric Oxide

PAGE Polyacrylamide Gel Electrophoresis

PARP poly(ADP-ribose) polymerase

PBS Phosphate Buffer Solution

PC2 Physical Containment Level 2

PCR Polymerase Chain Reaction

PES Phenazine Ethosulfate

PI Propidium Iodide

PI_DNA, -A, -H, -W Propidium Iodide, -Area, -Height, -Width

PIPES 1,4-piperazinediethanesulfonic acid

PMSF Phenylmethylsulfonyl fluoride

PS Phosphatidylserine

PSA Prostate Specific Antigen

Page 18: Investigating the effects of, the isoflavone, Phenoxodiol

xviii

PXD Phenoxodiol

QPCR Quantitative Polymerase Chain Reaction

RCF Relative Centrifugal Force = 0.00001118rN2

Redox Reduction Oxidation

RIPA Radioimmunoprecipitation

RNA Ribonucleic Acid

ROS Reactive Oxygen Species

RT PCR Reverse Transcriptase Polymerase Chain Reaction

SDS Sodium Dodecyl Sulfate

SDS-PAGE Sodium Dodecyl Sulfate Polyacrylamide Gel Electrophoresis

shRNA Short Hairpin Ribonucleic Acid

SNP Sodium Nitroprusside

SOD Superoxide Dismutase

SSC, -A, -H Side Scatter, -Area, -Height

TAE Tris-acetate EDTA

TBS-T Tris Buffered Saline with Tween

TCF T-cell factor

TE Tris EDTA

TEMED N,N,N',N'-Tetramethylethylenediamine

TNF-α Tumour Necrosis Factor Alpha

TRIS 2-Amino-2-(hydroxymethyl)-1,3-propanediol

TRIS-HCL 2-Amino-2-(hydroxymethyl)-1,3-propanediol with a pH set by

Hydrochloric acid addition

TRUS Transrectal ultrasound

TURP Transurethral Resection of Prostate

Versus Versus

xIAP X-linked Inhibitor of Apoptosis Protein

Z-DEVD-R110 Rhodamine 110 bis-(N-CBZ-L-aspartyl-L-glutamyl-L-valyl-L-aspartic

acid amide)

Z-VAD-FMK Carbobenzoxy-valyl-alanyl-aspartyl-(O-methyl)-

fluoromethylketone

Page 19: Investigating the effects of, the isoflavone, Phenoxodiol

xix

THESIS LAYOUT

This thesis comprises of 10 chapters divided into; Introduction, Literature Review,

Methodology, Four experimental chapters, General Discussion, Bibliography and Appendices.

Each experimental chapter is divided into; an introduction, brief methodology, results and

discussion section. The general discussion section is divided into discussion, conclusion and

limitations of the study.

ABSTRACT

In this study we investigated the cytotoxic effects of the isoflavone molecule Phenoxodiol, on

the prostate cancer cell lines LNCaP, DU145 and PC3. LNCaP cells represented a hormone

responsive early stage prostate cancer while DU145 and PC3 cells represented late stage

hormone refractory and chemoresistant prostate cancer. We investigated the cytotoxic effects

induced in these cell lines over a period of 24 or 48 hours of treatment with 10µM and 30µM

doses of phenoxodiol. Cells exhibited significant cytotoxicity and mitochondrial depolarisation

in response to phenoxodiol treatment. The study established LNCaP and DU145 cells

responded apoptotically to treatment, while PC3 cells respond necrotically. After confirming

cytotoxic responses the underlying signalling mechanisms were investigated which revealed

that phenoxodiol was not inducing cell death via an increase in Caspase-3 activity, in two of

three cell lines, and the apoptotic machinery of the cells responded in a varied manner

between cell lines. Phenoxodiol was revealed to work via Caspase independent actions and the

effect of the mitochondrial signalling molecules was not consistently altered in response to

treatment indicating phenoxodiol did not directly target intrinsic or extrinsic caspase signalling

pathways. We then investigated the effects of phenoxodiol on the cell cycle where treatment

groups exhibited a significant increase in G1 and S phase populations and a corresponding

decrease in G2 phase populations. The underlying signalling was investigated and it was

determined that p21WAF1 expression was increased consistently between the three cell lines in

Page 20: Investigating the effects of, the isoflavone, Phenoxodiol

xx

response to treatment. Finally we investigated the effects of phenoxodiol in combination with,

the G2 phase inhibitor, docetaxel and purified sFRP4 protein, a known agonist of the Wnt/β-

catenin growth regulation pathway. Phenoxodiol exhibited the ability to interfere with

docetaxel treatment as it prevented cells from reaching G2 phase, where docetaxel is

effective. Phenoxodiol exhibited the ability to increase the effectiveness of sFRP4 with

combination therapy significantly increasing the effectiveness of the two compounds

compared to individual responses. Phenoxodiol exhibits significant cytotoxicity, inducing

caspase independent cell death in the prostate cancer cell lines LNCaP, DU145 and PC3. All

three cell lines had significantly decreased viable cell populations, after only 48 hours of

treatment, through G1 and S phase cell cycle arrest. By targeting non classical apoptotic

pathways and successfully inducing cell cytotoxicity even in the most chemoresistant cell lines

and coupled with the reported ability of high tolerance of orally ingested phenoxodiol and few

reported side effects; Phenoxodiol represents a strong, effective, potential treatment for all

stages of prostate cancer.

Page 21: Investigating the effects of, the isoflavone, Phenoxodiol

1

1 . C H A P TER ONE: INT R OD UC TION

Prostatic adenocarcinoma is the second most commonly diagnosed malignancy in men and is

the most common cause of death in men in many western countries. Like other cancers,

prostate cancer has numerous clinical states ranging from a hormone-naïve clinically localised

primary tumour to lethal androgen-independent metastases (Arnold and Isaacs 2002).

Regulation of prostate growth is mediated via androgens and the corresponding androgen

receptor (AR) which regulates the transcription of target survival and apoptosis genes. Late-

phase metastatic disease is often androgen independent arising from either increased AR

expression and enhanced nuclear localisation of the AR, AR mutation resulting in a more

promiscuous receptor or the presence of alternative survival pathways (Bcl-2 upregulation)

that circumvent the AR (Feldman and Feldman 2001; Litvinov et al. 2003).

Epidemiological studies have consistently shown an inverse association between isoflavone

intake and the risk of cancer (Beecher 2003; Brown et al. 2005). In vitro mechanistic studies on

isoflavones provide insight into potential modes of anti-cancer action ranging from cell cycle

arrest and apoptosis induction to anti-angiogenic and anti-proliferative effects (Middleton et

al. 2000; Brown et al. 2005). Phenoxodiol, 2H-1-benzopyran-7-0,1,3-(4-hydroxyphenyl), is an

isoflavone derivative that has been shown to elicit cytotoxic effects against a broad range of

human cancers (Aguero et al. 2005; Axanova et al. 2005; Brown et al. 2005; Silasi et al. 2009).

Previous reports have indicated that phenoxodiol is a topoisomerase II inhibitor, inhibits

sphingosine kinase activity, downregulates transcription of angiogenic matrix metalloprotease

2 and other markers of angiogenesis, and can also induce apoptosis in chemoresistant ovarian

cancer cells by regulating anti-apoptotic signalling pathways (Constantinou et al. 2003;

Kamsteeg et al. 2003; Sapi et al. 2004; Aguero et al. 2005; Axanova et al. 2005; Brown et al.

2005; Gamble et al. 2005; Alvero et al. 2006). Phenoxodiol has been reported to inhibit cell

Page 22: Investigating the effects of, the isoflavone, Phenoxodiol

2

proliferation in a wide range of human cancer cell lines and induce G1 cell cycle arrest as

opposed to the G2-M arrest seen in similar flavanoid compounds like genistein (Alhasan et al.

1999; Aguero et al. 2005). It is postulated that phenoxodiol elicits its global anticancer activity

by modulating the sphingomyelin pathway (De Luca et al. 2005; Morre et al. 2007). Recent

data, however, have demonstrated that phenoxodiol is able to enhance perforin induced cell

death elicited by T-cells in both in vitro and in vivo models of colon cancer (Georgaki et al.

2009). Recent evidence has demonstrated that phenoxodiol augments the anti-cancer activity

of cisplatin against the DU145 prostate cancer cell line both in in vitro and in vivo studies. An

intracellular cisplatin accumulation assay showed a 35% (p<0.05) increase in the uptake of

cisplatin when cells were treated with a combination of 1μM cisplatin and 5μM phenoxodiol,

resulting in a 300% (p<0.05) increase in DNA adducts, hence explaining the sensitisation effect

(McPherson et al. 2009).

The mainstay of primary prostate treatment options is combined androgen ablation and

taxane therapy; while this approach is curative in early stage cases late stage treatment often

results in refractory disease development (Montero et al. 2005). Treatment options for

patients with late stage metastatic prostate carcinoma rely on the premise that androgen-

insensitive prostate carcinoma cells retain their basic cellular apoptotic machinery to undergo

programmed cell death, however increased tumour-cell resistance to apoptosis is an

underlying molecular reason contributing to disease progression and chemo-resistance

(Gleave et al. 2005). Apoptosis, the process of physiological cell death, is a focal area for the

study of cancer cell death. Apoptotic signalling is a mechanism of ordered cell death used by

the body to remove cells with damaged genetic material, and as such has evolved into a highly

complex and multi-channelled pathway to prevent accidental cell death from occurring.

Cancerous cells are able to prevent these messages of apoptosis from activating their

intracellular pro-apoptotic pathways. The ability of cancer cells to evade apoptosis is an

Page 23: Investigating the effects of, the isoflavone, Phenoxodiol

3

essential hallmark of cancer, and a key objective of cancer therapy is to restore cellular

sensitivity to apoptosis (Hanahan and Weinberg 2000).

In this study we investigated and quantified the cytotoxic effects of phenoxodiol treatment in

the prostate cancer cell lines LNCaP, DU145 and PC3 which represent early through late stage

prostate cancer. We investigated the type of cytotoxic response, the underlying cell death

signalling machinery, the effects on the cell cycle and the cell cycle signalling and the combined

effect of phenoxodiol and docetaxel and phenoxodiol and sFRP4, a Wnt/β-catenin cell growth

pathway agonist.

Page 24: Investigating the effects of, the isoflavone, Phenoxodiol

4

2 . C H A P TER T W O: L I TER A TURE REV IEW

2.1. CANCER

Cancer describes a range of diseases in which abnormal cells proliferate and spread out of

control. Common descriptions for cancer types are; benign or unlikely to harm the host,

malignant or harmful to the host and metastatic where a malignant tumour has spread into

other tissues in the body and implanted, growing in size. Other terms for cancer include

tumours and neoplasms, although these terms can also be used as descriptive terms for non-

cancerous growths. Carcinomas are malignant tumours arising from an epithelial origin while

sarcomas are malignant tumours arising from a connective tissue origin (Stevens 2009).

2.1.1. COMMON CANCERS AND PROSTATE CANCER IN MALES IN

AUSTRALIA

In 2007, there were more than 62,000 new cases of cancer diagnosed in males, which when

age-standardised equated to about 595 cases per 100,000 males. While age-standardised rates

of all cancers combined for males have risen since 1998, it is expected that they will steady at

about 568 new cases per 100,000 males between 2011 and 2020. This is primarily due to the

anticipated steadying of rates of prostate cancer incidence, coupled with decreasing rates in

lung cancer in males, which accounts for a further 9–10% of cases. When taking into account

expected changes in the population structure, this will translate into about 85,000 new cases

in males expected to be diagnosed in 2020 (Begg 2007).

Page 25: Investigating the effects of, the isoflavone, Phenoxodiol

5

TABLE 1: PREDICTED PROSTATE CANCER RATES 2011-2020, ADAPTED FROM BEGG, 2007

Year 0–24 years 25–44 years 45–64 years 65–84 years 85+ years

Count Rate Count Rate Count Rate Count Rate Count Rate

2011 35 0.9 30 0.9 7,750 276.8 11,460 891.6 1,660 1131.1

2012 35 0.9 30 0.9 8,060 286.1 12,150 902.6 1,750 1130.8

2013 35 1.0 30 1.0 8,420 296.3 12,790 913.8 1,850 1130.7

2014 40 1.0 35 1.0 8,820 306.6 13,400 923.2 1,930 1130.3

2015 40 1.0 35 1.0 9,210 316.5 14,030 932.3 2,020 1129.2

2016 40 1.1 35 1.1 9,640 326.3 14,660 940.9 2,100 1128.2

2017 45 1.1 40 1.1 10,070 336.3 15,250 945.2 2,160 1127.1

2018 45 1.2 40 1.2 10,490 346.6 15,890 951.2 2,220 1126.2

2019 45 1.2 40 1.2 10,930 357.7 16,550 958.9 2,280 1125.3

2020 50 1.2 45 1.2 11,350 368.6 17,220 966.5 2,340 1124.4

TABLE 2 THE 10 MOST COMMONLY OCCURING CANCER IN AUSTRALIA 2007, ADAPTED FROM BEGG

2007

Males Females

Site/type Cases ASR CI (95%) Site/type Cases ASR CI (95%)

Prostate 19,403 182.9 180.3–185.5

Breast 12,567 109.2 107.3–111.1

Bowel 7,804 75.2 73.5–76.9 Bowel 6,430 53.4 52.1–54.7

Melanoma of skin

5,980 57.2 55.7–58.7 Melanoma of skin

4,362 38.2 37.1–39.4

Lung 5,948 57.9 56.5–59.4 Lung 3,755 31.3 30.3–32.4

Lymphoid cancers

4,116 39.6 38.4–40.8 Lymphoid cancers

3,160 26.8 25.9–27.8

Myeloid cancers

1,859 18.5 17.7–19.4 Uterus 1,942 16.5 15.8–17.3

Kidney 1,716 16.3 15.5–17.1 Unknown primary

1,401 11.0 10.4–11.6

Bladder 1,644 16.5 15.7–17.3 Thyroid 1,331 12.2 11.6–12.9

Unknown primary

1,496 14.9 14.2–15.7 Ovary 1,266 10.8 10.2–11.4

Pancreas 1,352 13.1 12.4–13.8 Myeloid cancers

1,232 10.1 9.5–10.7

All cancers 62,019 595.1 599.8 All cancers 46,349 393.9 397.5

ASR: The rates were standardised to the Australian population as at 30 June 2001 and are

expressed per 100,000 population.

Page 26: Investigating the effects of, the isoflavone, Phenoxodiol

6

Prostate cancer accounts for about 30% of all new cases of cancer diagnosed in males

(excluding non-melanoma skin cancers), and is the second most common cause of cancer-

related death in males after lung cancer. With 19,403 new cases diagnosed in Australia in

2007, prostate cancer is a major public health concern (Begg 2007).

2.1.2. CANCER AS A DISEASE IN AUSTRALIA

In 2003, Cancer over took cardiovascular disease, 19% to 18%, as the biggest burden of disease

and injury in Australia. Four Fifths of that burden was from premature deaths, with Lung,

Colorectal, Breast and Prostate Cancer accounting for the leading majority of burden (Begg

2007).

FIGURE 1: BURDEN OF DISEASE IN AUSTRALIA IN 2003

Burden of disability adjusted life years (DALYs) by broad cause group expressed as: proportions

of total, proportions by sex and proportions due to fatal and non-fatal outcomes, Australia,

2003 (Begg 2007)

Page 27: Investigating the effects of, the isoflavone, Phenoxodiol

7

Total costs to the Australian health system were 2.9 billion or 5.8% of the total health

expenditure for 2003 (Begg 2007). Since 2003 statistics were released however, many newer,

high cost drugs have entered onto the market, examples of which are Herceptin and Avastin.

Their use for common cancers like breast and colorectal, will significantly increase the costs of

conventional cancer treatment as part of the Australian health expenditure.

2.2. PHENOXODIOL

Phenoxodiol ,[2H-1-benzopyran-7-0, 1,3-(4-hydroxyphenyl)], is a synthetic isoflavone molecule

first isolated from soy beans and now currently undergoing Phase III clinical trials for the

treatment of platinum and taxane refractory ovarian cancer (Morre et al. 2009). Phenoxodiol

has been shown to elicit cytotoxic effects against a broad range of human cancers (Sapi et al.

2004; Aguero et al. 2005; Axanova et al. 2005; Brown et al. 2005; Silasi et al. 2009; Aguero et

al. 2010). An isoflavone derivative, phenoxodiol has been shown to have a greater bio-

availability and increased potency/efficacy than its parent compound, which suffered in clinical

trials due to low bio availability (Aguero et al. 2005; Klein et al. 2007). Regardless of

progression in clinical trials, the primary cellular signalling target or targets of phenoxodiol

remain elusive and multiple studies covering various cancer cell types have detected different

methods of action (Straszewski-Chavez et al. 2004; Alvero et al. 2007; Georgaki et al. 2009;

Herst et al. 2009; McPherson et al. 2009; Saif et al. 2009; Aguero et al. 2010; de Souza et al.

2010; Wu et al. 2011).

Page 28: Investigating the effects of, the isoflavone, Phenoxodiol

8

FIGURE 2 PHENOXODIOL MOLECULAR STRUCTURE

The molecular structure of Phenoxodiol represented in 3D as presented at www.novogen.com

2.2.1. FLAVANOIDS AND ISOFLAVONES

In vitro mechanistic studies on isoflavones provide insight into potential modes of anti-cancer

action ranging from cell cycle arrest and apoptosis induction to anti-angiogenic and anti-

proliferative effects (Middleton et al. 2000; Brown et al. 2005). Strong in vitro evidence exists

for the activity of a variety of isoflavones on hormone sensitive and insensitive prostate cancer

cell lines (Hempstock et al. 1998; Mitchell et al. 2000; Hedlund et al. 2003) and in vivo in rats

(Risbridger et al. 2001). Isoflavones appear to have pleiotropic effects on prostate cancer cells,

including an ability to exert hormonal influences. Phenoxodiol is a synthetic isoflavone

metabolite that is a natural intermediate (dehydroequol, 7,40-dihydroxyisoflav-3-ene) in the

metabolism of daidzein to equol (Joannou et al. 1995). It is cytotoxic in vitro and in vivo in rats

(Constantinou et al. 2003; Mor et al. 2006). It may be capable of re-sensitising platinum- and

taxane-resistant ovarian cancer cells in vitro (Kamsteeg et al. 2003; Sapi et al. 2004; Kluger et

al. 2007) and appears to have antiangiogenic (Gamble et al. 2005) and anti-inflammatory

Page 29: Investigating the effects of, the isoflavone, Phenoxodiol

9

properties (Widyarini et al. 2001). It has improved bioavailability when compared with

genistein and low toxicity in clinical trials (Kelly and Husband 2003; Davies 2005; de Souza et al.

2006).

FIGURE 3 PRODUCTION OF EQUOL FROM DAIDZEIN

Production of Phenoxodiol naturally occurs as a process of Daidzein to Equol formation

(Joannou et al. 1995). Figure adapted from (Jian 2009).

Page 30: Investigating the effects of, the isoflavone, Phenoxodiol

10

2.2.2. PHENOXODIOLS REPORTED METHOD OF ACTION

Phenoxodiol has been reported to elicit diverse cytotoxicity activity via several mechanisms

including induction of G1 arrest via p53-independent p21WAF1 regulation, resulting in loss of

cyclin-D kinase activity in HN12 head and neck cancer cells, strong S-phase arrest in ovarian

cancer cells at high concentrations, restoration of death receptor-refractory cancer cells to

respond to extracellular death signals by re-enabling the transduction of an activated Fas or

TRAIL receptor signal to Caspase-8 due to phenoxodiol induced degradation of Flice Like

Inhibitory Proteins (FLIP) which enables activation of the extrinsic apoptosis pathway (Aguero

et al. 2005; Alvero et al. 2006; Alvero et al. 2007; Alvero et al. 2008; Aguero et al. 2010).

Phenoxodiol also engages the intrinsic apoptosis pathway, which is typically non-functional in

late stage cancer due to overexpression of pro-survival factors such as sphingosine-1

phosphate, which drives the stabilisation of Akt causing an accumulation of XIAP and

subsequent inhibition of Caspase-3 and -9 (Choueiri et al. 2006; Alvero et al. 2008). Recent

data, however, have demonstrated that phenoxodiol is able to enhance perforin induced cell

death elicited by T-cells in both in vitro and in vivo models of colon cancer (Georgaki et al.

2009). Recent evidence has demonstrated that phenoxodiol augments the anti-cancer activity

of cisplatin against the DU145 prostate cancer cell line both in in vitro and in vivo studies. An

intracellular cisplatin accumulation assay showed a 35% (p<0.05) increase in the uptake of

cisplatin when cells were treated with a combination of 1μM cisplatin and 5μM phenoxodiol,

resulting in a 300% (p<0.05) increase in DNA adducts, hence explaining the sensitisation effect

(McPherson et al. 2009).

An early event in the phenoxodiol mechanism of action is the disruption of the

sphingomyelinase pathway resulting in an accumulation of ceramide, which drives the

truncation of Bid by Caspase-2, thereby resulting in Bid translocation to the mitochondria.

Caspase-8 also contributes to Bid formation accompanied by Bax translocation to the

Page 31: Investigating the effects of, the isoflavone, Phenoxodiol

11

mitochondria, both Bax and Bid serve to depolarise the mitochondria, resulting in Cytochrome

c release and the formation of the apoptosome utilising Caspase 9. Key to the mechanism by

which phenoxodiol enables the reactivation of caspase-mediated apoptosis is its ability to

promote X Linked Inhibitor of Apoptosis Protein (XIAP) removal. This is achieved by the release

of SMAC-Diablo and OMI-Htra2 from the mitochondria, which both serve to reduce the

intracellular content of XIAP, thereby allowing the full activation of Caspase-3 and -9. XIAP was

also found to be targeted for degradation by the proteasome in phenoxodiol treated cells,

which also further explains the active removal of intracellular XIAP (Saif et al. 2009). Active

XIAP removal sheds light on the mechanism by which phenoxodiol acts as a chemosensitiser in

refractory ovarian cancer and melanoma as expression of XIAP is linked with chemoresistance

(Kamsteeg et al. 2003; Mahoney 2007).

Phenoxodiol is reported to be a topoisomerase II inhibitor, inhibiting sphingosine kinase

activity, downregulating transcription of angiogenic matrix metalloprotease 2 and other

markers of angiogenesis and inducing apoptosis in chemoresistant ovarian cancer cells by

regulating anti-apoptotic signalling pathways (Constantinou et al. 2003; Kamsteeg et al. 2003;

Aguero et al. 2005; Axanova et al. 2005; Brown et al. 2005; Gamble et al. 2005; Alvero et al.

2006). Phenoxodiol has exhibited the ability to target a subset of NADH oxidases (NOX) that is

thought to be a terminal oxidase primarily expressed on cancer cells (Herst et al. 2007). These

cell surface NADH oxidases form part of a family of ECTO-NOX proteins that play an important

role in red-ox regulation (Yagiz et al. 2007). Phenoxodiol has been reported to inhibit cell

proliferation in a wide range of human cancer cell lines and induce G1 cell cycle arrest as

opposed to the G2-M arrest seen in similar flavanoid compounds like genistein (Alhasan et al.

1999; Aguero et al. 2005). It is postulated that phenoxodiol elicits its global anticancer activity

by modulating the sphingomyelin pathway (De Luca et al. 2005; Moore et al. 2007).

Page 32: Investigating the effects of, the isoflavone, Phenoxodiol

12

Phenoxodiol has been demonstrated to successfully induce cytotoxicity across multiple cancer

cell types, through multiple and varied reported methods of action.

FIGURE 4 PHENOXODIOLS REPORTED METHOD OF ACTION AS ADAPTED FROM SILASI ET AL. 2009

Overexpression of AKT, XIAP and FLIP are linked with cancer progression. XIAP inhibits both

death receptor and mitochondrial-mediated apoptosis, while FLIP predominantly inhibits death

receptor apoptosis. Phenoxodiol initiates a cascade of intracellular events including ceramide

accumulation, caspase 2 activation, and inhibition of sphingosine kinase activity, which results

in reduced intracellularsphingosine-1-phosphate (S1P). Reduced S1P disrupts AKT-p formation

and stabilisation of the AKT protein, thereby promoting AKT, XIAP and FLIP total protein

degradation. Mitochondrial depolarisation results in the release of the serine protease

Omi/HtrA2, which contributes to caspase-mediated cleavage of XIAP and Smac-Diablo which

sequesters and inhibits XIAP. Concomitant with Omi/Hrta2 and Smac-Diablo appearance in the

cytosol, Cytochrome c is also released from the mitochondria thereby causing the activation of

Caspase-9 and subsequent activation of Caspase-3. Due to FLIPs removal, Caspase-8 is also

activated contributing to Caspase-3 activation via the mitochondria. Adapted from (Silasi et al.

2009)

Page 33: Investigating the effects of, the isoflavone, Phenoxodiol

13

2.2.3. SYNERGISTIC ACTIVTY

Synergy with other chemotherapeutic agents has been described using mouse xenograft

models, Alvero and colleagues demonstrated significant synergy of phenoxodiol in

combination with carboplatin (decrease in tumour mass from 6% with carboplatin as

monotherapy to 47% in combination) and paclitaxel (35 – 74%), as well as with gemcitabine

(Alvero et al. 2007; Alvero et al. 2008). Pretreatment with phenoxodiol was found to shift the

IC 50 of each of these drugs downward while co-administration with topotecan also decreased

the IC 50 and have efficacy in 5/9 ovarian cell lines, which were known to be resistant to

topotecan monotherapy (Alvero et al. 2007). By using melanoma microarrays and automated

quantitative analysis technology, (Kluger et al. 2007) demonstrated that pretreatment with

and a novel agent, triphendiol, for the treatment of pancreaticobiliary cancers sensitise

platinum resistant melanoma cells to carboplatin. Moreover, they linked this effect to the

increased cleavage and degradation of XIAP, a direct inhibitor of Caspase-9, -3 and -7 that

leads to decreased initiation and execution of apoptosis and has long been associated with

resistance to chemotherapy in a number of malignancies.

2.2.4. TREATMENT WITH PHENOXODIOL

Phenoxodiol is rapidly absorbed after oral administration and maximum plasma concentration

occurs after 3 hours. In the plasma, phenoxodiol is present almost exclusively (99%) in a

conjugated state with glucuronide and/or sulfate moieties and is highly bound to human

plasma proteins (80 – 95%) with the greatest affinity for albumin. It binds to a lesser extent to

α-1-acid glycoprotein and sex-hormone-binding globulin. The half-life of the drug is 8 – 10 h

and excretion is predominantly urinary in its conjugated form (Silasi et al. 2009). Following

continuous intravenous infusion, plasma concentrations of free plasma phenoxodiol rose

rapidly with an apparent accumulation half-life of 0.17 hours (Howes et al. 2011). Due to the

Page 34: Investigating the effects of, the isoflavone, Phenoxodiol

14

high clearance rate of phenoxodiol it would appear that administration by continuous infusion

or by chronic oral administration may be the optimal modes of administration if it considered

that constant plasma levels are desirable for anti-cancer therapy (Howes et al. 2011).

Following multiple Phase I, II trials it appears that phenoxodiol is a multi-pathway initiator of

apoptosis with broad anti-tumour activity and high specificity for tumour cells and although

clinical trials are still ongoing, phenoxodiol appears to be particularly suited for reversal of

chemo-resistance and its activity is being investigated as a chemo-sensitiser of standard

chemotherapy agents in solid cancers. Phenoxodiol was granted ‘fast track’ status by the US

FDA in its development as a chemosensitiser for platinum and taxane drugs used in the

treatment of recurrent ovarian cancer (Silasi et al. 2009).

2.3. CARCINOGENESIS / CANCER DEVELOPMENT

Carcinogenesis is the development of cancerous tissue from single cells. It is a multistage

process that assumes cancer is a clonal development from a single cell that has been blocked

from terminal differentiation. Growth signal autonomy, insensitivity to antigrowth signals and

resistance to apoptosis all lead to an uncoupling of the cell’s growth program from signals in

the environment. In 2000, a paper titled The Hallmarks of Cancer condensed and abridged the

common pathways that cancer cells must progress through to develop from benign to

malignant and finally metastatic tumour phenotypes (Hanahan and Weinberg 2000). There

were six potential characteristics cells could acquire to develop into cancerous cells.

1) Evading apoptosis

2) Self-sufficiency in growth signals

3) Insensitivity to anti-growth signals

4) Limitless replicative potential

5) Sustained angiogenesis

*Bolded items were investigated in this

thesis.

Page 35: Investigating the effects of, the isoflavone, Phenoxodiol

15

6) Tissue invasion of metastasis

In 2004 (Trosko et al. 2004) argued the previous six phases were important phenotypic

markers, as well as concepts, but the role of stem cells and cell-cell communication was

equally important in determining cancer development.

7) Pluripotent stem cell differentiation and development

8) Gap junctional intercellular connection

(Trosko et al. 2004) states that cancer cell development must go through initiator, promotion

and finally progression phases. Stem cell development is involved in the initiator phase. The

promotion phase covers potentially reversible or interruptible clonal expansion of the initiated

cell by a combination of growth stimulation and inhibition of apoptosis. When the expanded

initiated cells accrue sufficient mutations and epigenetic alterations to become growth

stimulus independent and resistant to growth inhibitors and apoptosis, to have unlimited

replicative potential and invasive and metastatic phenotypes, then the progression phase has

been achieved. The promotion and progression phase is targeted by conventional cancer

treatment.

Page 36: Investigating the effects of, the isoflavone, Phenoxodiol

16

FIGURE 5: ACQUIRED CAPABILITIES OF CANCER ADAPTED FROM HANAHAN ET AL 2000 AND TROSKO ET

AL 2004

The eight potential characteristics that cancer cells can develop, or develop from, as referred to

in Hanahan et al. 2000 and Trosko et al. 2004.

Page 37: Investigating the effects of, the isoflavone, Phenoxodiol

17

2.3.1. EVASION OF APOPTOSIS

Apoptosis is the programmed physiological cell death of a cell in response to both extrinsic and

intrinsic factors. The ability for prostate cancer cells to avoid apoptotic signalling is considered

fundamental in cancer development. The capacity for a population of cells to expand is not

controlled just by growth rate, but also by attrition rate, and attrition is controlled by

apoptosis (Hanahan and Weinberg 2000). Evasion of apoptosis was one of the first

characteristics of cancer discovered when cells began to no longer typically respond to

chemotherapy treatments as they once did. The evidence from mouse models, cultured cells

and biopsies is suggesting that most, if not all cancers have the ability to evade a majority of

apoptotic signalling. Apoptosis has also been the focus of targeted drug design. The ability to

resensitise tumours to apoptotic signalling is considered critical in modern chemotherapy

design as it would make tumour masses far easier to treat. It would also allow the use of older

drugs, in combination potentially lowering doses needed resulting in reduced side effects,

lower mortality and hopefully decreased recurrence rates (Brown and Wouters 1999; Brown

and Wilson 2003; Brown and Attardi 2005).

2.3.2. SELF-SUFFICIENCY IN GROWTH SIGNALS

Normal cells require mitogenic growth signals to enter into a proliferative state from a

quiescent state, cancer cells exhibit an opposite behaviour with tumour cells generating many

of their own growth signals, reducing their dependence on the stimulation of the normal tissue

environment (Hanahan and Weinberg 2000). Prostate cells have a particular dependence on

the hormone testosterone and are stimulated by the androgen receptor (AR) to proliferate.

The transition to androgen independence is a multifaceted process that involves selection and

outgrowth of cells less dependent on androgen stimulation as well as adaptive upregulation of

genes that help the cancer cells survive and grow after androgen ablation (So et al. 2003).

Page 38: Investigating the effects of, the isoflavone, Phenoxodiol

18

Prostate tumours are composed of various subpopulations of cells that respond differently to

androgen withdrawal therapy and the development of a mutated AR or null AR is characteristic

of late stage prostate cancer where the cells have gained self-sufficiency in growth signals

(Arnold and Isaacs 2002). One signalling pathway that has been implicated in prostate

progression and late stage independence from androgen stimulation is the Wnt family of

proteins. A key pathway of cellular homeostasis and proliferation, aberrant Wnt signalling

activates ⁄ stimulates proteins and respective target genes, which drive prostate cancer

progression. Wnt5a has been implicated in aggressive metastasis and prostate cancer relapse

after prostatectomy and the cell surface frizzled receptors are overexpressed in prostate

cancer. However this overexpression is counterbalanced by the secreted Frizzled Related

Protein family (sFRP), which interact with Frizzled receptors expressed in prostate cancer cells,

and this heterodimer suppresses AR-mediated transactivation (Farooqi et al. 2011).

2.3.3. INSENSITIVITY TO ANTI-GROWTH SIGNALS

In conjunction with insensitivity to apoptosis signalling; prostate cancer cells develop

insensitivity to anti-growth signals. Within a normal tissue, multiple antiproliferative signals

exist to maintain cellular quiescence and tissue homeostasis, including soluble extracellular

growth inhibitory signals and intracellular signals (Hanahan and Weinberg 2000). Anti-growth

signals typically force cells into a quiescent state or further differentiated, post-mitotic cell

type and can affect the expression of integrins and other cell adhesion modelcules, the result

preventing cells from moving past the G1 phase of the cell cycle by affecting Cyclin-D1

expression (Ladha et al. 1998). Avoidance of anti-growth signals is not enough to solely cause

progression of cancer, a reliance on c-Myc or the Wnt/β-catenin pathway in stimulating

mitogenic activity is considered an integral part of the development of insensitivity to anti-

growth signals through stimulation of pro-growth pathways (He et al. 1998).

Page 39: Investigating the effects of, the isoflavone, Phenoxodiol

19

2.3.4. LIMITLESS REPLICATIVE POTENTIAL

Even if cells have developed self-sufficiency in growth signals and an ability to evade apoptosis

they are still limited in mitogenic capability by normal cell replication machinery. The Hayflick

limit (or Hayflick Phenomenon) is the number of times a normal cell population will divide

before it stops, presumably because the telomeres shorten to a critical length. Mammalian

cells in a cell culture divide between 40 and 60 times then enter a senescence phase with each

mitotic division effectively shortening the telomeres on the DNA of the cell (Hayflick and

Moorhead 1961). Telomere shortening in humans eventually makes cell division impossible,

and it is presumed to correlate. Maintenance of the length of the telomeric region appears to

prevent genomic instability and the development of cancer. Upon activation of mitogenic

signalling, cells commit to entry into a series of regulated steps allowing completion of the cell

cycle. Cells begin in G1 phase, the time between M and S phases, and before entry into S

phase, where DNA is replicated, must pass through a restriction point (Pardee 1974) that

analyses and attempts to repair DNA damage. After S phase, cells enter G2 phase (the time

between the S and M phases) where cells can repair errors that occurred during DNA

duplication, preventing the propagation of these errors to daughter cells. Finally, the

separation into two daughter cells by chromatid separation occurs and is called M phase

(Senderowicz 2004). The sequence of events in cell cycle progression is highly orchestrated and

depends on the cyclic activation and inactivation of cyclin dependent kinases (CDK), which

govern the progression of the cells from one phase to another. In the event of tumourigenesis,

constitutive mitogenic signalling as well as mutations in tumour suppressor genes and proto-

oncogenes leads to cell cycle deregulation and uncontrolled proliferation (MacLachlan et al.

1995).

Page 40: Investigating the effects of, the isoflavone, Phenoxodiol

20

2.3.5. SUSTAINED ANGIOGENESIS

The growth of new blood vessels, angiogenesis, is critical in the development of the cancer

mass (Berry and Eisenberger 2005; Alvero et al. 2008). Solid tumour masses require large

amounts of energy to continue to replicate and grow at such rapid rates, the oxygen and

nutrients consumed by these cells is crucial and results in nearly all cells residing within 100µm

of a capillary blood vessel (Hanahan and Weinberg 2000). The development of blood vessels in

normal tissue and organs occurs at organogenesis in the foetal life, in comparison angiogenesis

occurs whenever irregular vascularisation signalling occurs, called the angiogenic switch

(Carmeliet and Jain 2011). Tumour-related angiogenesis supports tumour growth and is also a

major obstacle for successful immune therapy as it prevents migration of immune effector

cells into established tumour parenchyma (Hamzah et al. 2008). Thus angiogenesis can protect

the cancer mass from the hosts immune system while providing it with energy and oxygen. The

coordinated expression of pro- and antiangiogenic signalling molecules, and their modulation

by proteolysis, appear to reflect the complex homeostatic regulation of normal tissue

angiogenesis and of vascular integrity (Avraamides et al. 2008).

2.3.6. TISSUE INVASION AND METASTASIS

Tissue invasion and metastasis of cancer relies upon cells having developed the ability to evade

apoptosis as well as ignore the lack of gap junctional responses. Anoikis, the ability of cells to

evade apoptosis when shed from the primary tumour mass, is a key component in metastatic

invasion of tissue (Langley and Fidler 2011). Invasion relies on the angiogenic ability of the cells

and is considered one of the last Approximately one in eight prostate cancer cases

metastasises widely, typically to bone, adjacent soft tissue, liver, and lung and is consistent

with poor prognosis and outcome (Kleeberger et al. 2007). Development of a metastatic

phenotype is the result of an accumulation of many gene dysregulations and is considered a

Page 41: Investigating the effects of, the isoflavone, Phenoxodiol

21

key step in the progression from benign to metastatic and malignant tumour phenotypes

(Hanahan and Weinberg 2000).

2.3.7. PLURIPOTENT STEM CELL DIFFERENTIATION AND

DEVELOPMENT

The cancer stem cell hypothesis has two central tenets—tumours are derived from

transformation of normal stem cells or their progeny (i.e., progenitor or even differentiated

cells) and every tumour contains a small population of stem-like cells that possess a unique

ability to drive tumour formation and maintain tumour homeostasis (Reya et al. 2001). Current

theory is that stem cells are immortal and only become mortal once they enter into a terminal

differentiation cycle, as seen in stem cell derived tumours that can be serially initiated in mice

including solid tumours (Li et al. 2008).

Thus, stem cells are a self-renewing, immortalised group of cells that have been determined to

exist in nearly all organs and tissues (Reya et al. 2001). Stem cell populations normally make up

to 0.01% of a total cell population, however in cancerous tissues stem cells have been found

from 0.1-0.2% of total cells, an increase of ten to twenty fold. This increase in self renewing

cells that are not easily targeted by conventional chemotherapy, due to their slow cell cycling,

receptor complexes and adaptive ability is one theory used to describe the ability of cancer to

become refractive over time to initially successful treatment (Li et al. 2008).

Page 42: Investigating the effects of, the isoflavone, Phenoxodiol

22

FIGURE 6: STEM CELL THEORY ADAPTED FROM TROSKO ET AL 2004

Showing the characteristics of stem cell theory, Figure 6 demonstrates how slow cycling stem

cells regenerate and differentiate normally. Following initiation stem cells can still differentiate

normally, however clonal expansion and promotion of initiated cells can result in progression

phase, where cancer cells develop malignancy.

2.3.8. GAP JUNCTIONAL INTERCELLULAR CONNECTIONS

Gap junctional intercellular connections (GIJC) are involved in carcinogenesis in various ways.

Most, if not all, tumour cells lack functional GJIC (Trosko et al. 2004). Cancer cells lack the

expression of connexion genes due to the activation of various oncogenes or the loss of

specific tumour suppressor genes, resulting in a cell less affected by the extracellular

environment and less likely to differentiate. Most, if not all, tumour promoting chemicals and

conditions reversibly inhibit GJIC and oncogenes, such as Src, Ras, Raf, have been reported to

down regulate GJIC. Antisense connexin genes transfected in normal cells induces a

tumourigenic phenotype while connexin genes transfected into tumour cells restore growth

control and reduce the tumourigenicity of the cells (King et al. 2000).

Page 43: Investigating the effects of, the isoflavone, Phenoxodiol

23

2.4. THE PROSTATE

The prostate is a walnut sized gland of the male accessory reproductive system that surrounds

the urethra and ejaculatory duct directly inferior to the urinary bladder (Saladin 2007). It

measures 2x4x3 cm and is an aggregate of 30-50 compound tubuloacinar glands enclosed in a

single fibrous capsule, these glands empty via ~20 pores into the urethral wall. The position of

the prostate immediately anterior to the rectum allows for it to be palpated through the rectal

wall for lumps suggestive of prostate cancer (Saladin 2007). The prostate is approximately ½

glandular, ¼ involuntary muscle, ¼ fibrous tissue and the structure has a dense fibrous capsule

of the prostate that incorporates the prostatic plexuses of nerves and veins (Basmajian and

Slonecker 1989; Moore et al. 2007). The prostate is surrounded by the visceral layer of the

pelvic fascia, forming a fibrous prostatic sheath that is thin anteriorly, continuous

anterolaterally with the puboprostatic ligaments, and dense posteriorly, continuous with the

rectovesical septum, surgically called the fascia of Denonvilliers which separates the prostate

and urinary bladder from the rectum (Basmajian and Slonecker 1989; Moore et al. 2007).

The prostate is a modified portion of the urethral with the glands organised in 3 concentric

groups; mucosal are short and simple and open all around the urethra while the intermediate

or submucosal glands open into the prostatic sinus. The outermost or main glands are long and

branching and envelope the other groups except in front where those of opposite sides are

joined by a nonglandular isthmus and nearly all hypertrophies of the prostate arise from these

mucusol and submucosal glands (Basmajian and Slonecker 1989).

The prostate arterial blood supply is predominantly from branches of the internal iliac artery

including the inferior vesical arteries, internal pudendal and middle rectal arteries (Moore et

al. 2007). The prostatic venous drainage is complicated with the deep dorsal vein of the penis

Page 44: Investigating the effects of, the isoflavone, Phenoxodiol

24

draining the front plexus, the vesical plexus draining superiorly and posteriorly draining into

the internal iliac veins while also potentially anastamosing with the vertebra plexus of veins

and all of this occurring between the fibrous capsule of the prostate and the prostatic sheath

(Basmajian and Slonecker 1989). Due to this varied blood drainage prostate cancer frequently

spreads to the central nervous system (CNS), lower lumbar vertebrae, pelvic bones and femora

(Basmajian and Slonecker 1989). The nervous innervations of the prostate are autonomic, with

parasympathetic efferent fibres arising from pelvic splanchnic nerves (sacral nerves 2, 3 and 4)

while sympathetic efferent fibres come from sacral splanchnic nerves (arising from

sympathetic trunk) and together they help form the inferior hypogastric plexus which supplies

the prostate gland. The lymphatic drainage of the prostate is chiefly into the internal iliac

nodes but some pass to the sacral lymph nodes with implications for metastasising prostate

cancers ability to spread into spinal cord and CNS (Moore et al. 2007).

The prostate gland is histologically divided into three major regions – the peripheral, central

and transitional zones. While most cancers develop in the peripheral zone (68%), fewer

originate in the transitional (24%) or central (8%) zone (McNeal et al. 1988). Each zone consists

of ducts and acini lined by an epithelial sheet. The epithelium consists of a bi-layer of basal

cells beneath the secretory, luminal cells and is interspersed with neuroendocrine cells. The

majority of the basal cells are androgen receptor (AR) negative and consist of self-renewing

stem cells that differentiate into transit amplifying cells (also AR negative) with limited

proliferative capacity. These transit amplifying cells clonally expand, differentiate and migrate

from the basal to the luminal layer where they differentiate to form mature, secretory luminal

cells that are nonproliferative and AR-positive (Litvinov et al. 2003; Uzgare et al. 2004). A

minority of basal stem cells differentiate into neuroendocrine cells that are AR-negative and

secrete specific peptides (Arnold and Isaacs 2002).

Page 45: Investigating the effects of, the isoflavone, Phenoxodiol

25

2.4.1. TESTING FOR PROSTATE CANCER

Common testing for prostate cancer involves the digital rectal exam (DRE), detection of benign

prostatic hyperplasia (BPH) and admission of the prostatic specific antigen (PSA) blood exam.

The position of the prostate surrounding the urethra means that, during conditions of benign

prostatic hyperplasia, a transurethral resection of the prostate (TURP) will be performed and

the tissue analysed for potential cancer pathology. The position of the prostate anterior to the

rectum allows for a trans-rectal ultrasound biopsy (TRUS) to be performed when high PSA

scores indicate a potential for the tissue to become cancerous, the TRUS technique allows for

ultrasound to guide the biopsy needle to the abnormal looking region (Park et al. 2012).

Following biopsy using the TURP or TRUS methods cells are stained and compared to the

Gleason scale, which compares how differentiated cells are, and gives a 1-5 grade (Gleason

1977). Grades between two separate regions are added and result in a Gleason score (also

known as sum or pattern) which results in a prognosis with a minimum score or 2 and a

maximum or 10, the higher the number the higher the chance the cancer has reached

malignant or metastatic phenotypes. The move from benign to metastatic prostate cancer is

complicated but is based around the development of the androgen independent,

undifferentiated and subsequent invasive phenotypes (Feldman and Feldman 2001; So et al.

2003).

2.4.2. PHENOTYPIC PROGRESSION OF PROSTATE CANCER

Prostate cancer progresses from a benign to malignant to metastatic phenotype in conjunction

with a loss of androgen receptor function and a gain of invasive potential (So et al. 2003). The

prostatic epithelium is surrounded by a fibromuscular stroma that contains AR-positive

smooth muscle cells and fibroblasts amongst other cell types. Androgen occupancy of nuclear

Page 46: Investigating the effects of, the isoflavone, Phenoxodiol

26

AR in these stromal cells results in their production and secretion of paracrine growth and

survival factors for the epithelium (Litvinov et al. 2003).

Transformation to a malignant phenotype involves a shift from this paracrine axis to a situation

in which AR signalling directly activates the production of autocrine growth and survival factors

by the prostate cancer cells themselves as well as regulates the production of secretory

proteins by these malignant cells. Thus androgen acts in an autocrine manner as the major

regulator of proliferation and survival in such androgen responsive prostate cancer cells

(Litvinov et al. 2003). It is for this reason that androgen blockade via administration of

leutenizing hormone releasing hormone (LHRH) analogues results in inhibition of tumour

growth and a positive clinical response. Cellular adaptation to low levels of circulating

androgen along with clonal expansion, however, results in the emergence of hormone

refractory tumours with or without metastases (Gao et al. 2001; Isaacs and Isaacs 2004).

Understanding the mechanisms for development of hormone refractory cancer (i.e.

dysregulation of the signalling network) requires knowledge of the cellular organisation and

the cell of origin for the cancer. Prostate cancer is very common in developed countries and is

widely variable in clinical course. Most cases remain confined to the prostate and adjacent soft

tissue and cause no harm. However, approximately one in eight cases metastasises widely,

typically to bone, adjacent soft tissue, liver, and lung (Kleeberger et al. 2007).

FIGURE 7 CHARACTERISTIC PROGRESSION OF PROSTATE CANCER FROM EARLY TO LATE STAGE

Page 47: Investigating the effects of, the isoflavone, Phenoxodiol

27

2.4.3. TREATMENT OF PROSTATE CANCER

Age and hormone are two known factors influencing the incidence of prostate cancer and,

because of that, cancer cells initially respond to androgen withdrawal by undergoing apoptosis

among the hormone-dependent population. However, patients with advanced or metastatic

prostate cancer develop hormone refractory status that becomes fatal because of the growth

of androgen-independent tumour cells and the emergence of tumour clones. Therefore, the

potential cancer chemotherapy to cause apoptosis in metastatic prostate cancer is necessary

and urgent for the clinical treatment (Hotte and Saad 2010; Parente et al. 2012).

Patients with localised disease are candidates for ionizing radiation therapy. Patients with

metastatic carcinoma of the prostate are usually treated with androgen ablation therapy. After

further disease progression, they may be treated with chemotherapy including docetaxel and

doxorubicin (Sweat 2005). Ionizing radiation and doxorubicin are DNA-damaging agents that

induce double-strand breaks in DNA. This leads to the activation ofDNA damage checkpoints.

One of the key proteins in these pathways is the tumour suppressor p53, which triggers cell

cycle arrest and induces the repair of DNA damage, responses needed for cell survival, or

alternatively, apoptosis (Devlin et al. 2008).

The interaction between testosterone and the androgen receptor (AR) is essential for prostate

development. Because AR signalling has also been shown to play a key role in prostate

carcinogenesis, androgen ablation therapy (AAT) is a commonly used form of treatment,

particularly for advanced disease. While AAT leads to significant levels of prostate cancer cell

Page 48: Investigating the effects of, the isoflavone, Phenoxodiol

28

apoptosis, the effect is short-lived and ultimately not curative as most patients develop

androgen-independent disease (D'Antonio et al. 2008).

2.4.4. CHEMOTHERAPY AND PROSTATE CANCER

Currently, only patients who have detectable macroscopic metastatic disease should receive

systemic chemotherapy outside of a clinical trial. As any treatment for advanced disease

remains palliative, patients with advanced prostate cancer are encouraged to participate in

clinical trials (Hotte and Saad 2010). Combined docetaxel (a taxane drug that induces

polymerisation of microtubules and phosphorylation of the Bcl-2 protein) and prednisone is

currently considered the standard of care for men with detectable metastatic disease, based

largely on the simultaneous publication of two large randomised controlled trials comparing

this combination with the previously established standard of mitoxantrone and prednisone

(Parente et al. 2012). (Petrylak et al. 2004; Tannock et al. 2004) simultaneously published

combined trials of docetaxel and prednisone. Men in both trials had clinical evidence of

metastases with or without symptoms and had undergone anti-androgen withdrawal

response. Overall survival was the primary endpoint in both trials. (Tannock et al. 2004)

reported improved survival with docetaxel (every-3-weeks dosing) compared with

mitoxantrone–prednisone, median survival: 18.9 months versus 16.5 months(Petrylak et al.

2004) reported longer survival time with docetaxel–estramustine combination chemotherapy

as compared with mitoxantrone, median survival: 17.5 months versus 15.6 months. Late stage

prostate cancer treatment still remains an attempt at extension of life and is not yet

considered curative (Hotte and Saad 2010).

Page 49: Investigating the effects of, the isoflavone, Phenoxodiol

29

FIGURE 8 DEVELOPMENT OF RESISTANCE TO CHEMOTHERAPEUTIC TREATMENT ADAPTED FROM

JOHNSTONE, RUEFLI AND LOWE 2002

Addition of chemotherapeutic drugs to tumour cells results in interaction between the drug and

intracellular targets, and induction of the primary effect. Classical drug resistance proteins such

as drug efflux pumps can inhibit the primary effect by preventing drug-target interactions.

Depending on the severity of the initial insult, drug-induced damage may result in cytotoxic cell

death or initiate a series of secondary effects mediated by various stress signalling pathways

leading to cell death or cell cycle arrest. Consequently, mutations in these downstream events

can produce multidrug resistance. Adapted from (Johnstone et al. 2002).

2.5. APOPTOSIS

In classic apoptosis cellular membranes are disrupted, the cytoplasmic and nuclear skeletons

broken down and nuclear material fragmented and within the space of 30 to 120 minutes the

cell is destroyed and engulfed by nearby cells with no inflammatory response (Wyllie et al.

Page 50: Investigating the effects of, the isoflavone, Phenoxodiol

30

1980). When first described in 1972 (Kerr et al. 1972) raised the possibility of apoptosis being a

barrier to cancer after discovering massive apoptosis in populations of rapidly growing

hormone dependent tumour cells following hormone withdrawal. Resistance to apoptosis has

been investigated intensely for the last 20 years with some characteristic pathway mutations

becoming recognised in many different forms of cancer.

The loss of a functional p53 oncogene was described in 1996 (Harris 1996) and is discovered in

well over 50% of all cancers. This causes the removal of one of the key DNA damage sensing

apparatus in the cell and causes significant inability to induce apoptosis in response to DNA

damage, allowing more errors to accumulate rapidly. Abnormalities, including hypoxia and

oncogene hyper-expression, are also funnelled in part via p53 to the apoptotic machinery;

these too are impaired at eliciting apoptotic activity when p53 function is lost (Levine 1997).

Discovery of the anti-apoptotic Bcl-2 gene (Korsmeyer 1992) and later its family of related

signalling molecules further significantly enhanced the knowledge of how cancer cells develop

resistance to apoptotic stimulation.

The ability for cancer cells to ignore anoikis, the process of apoptosis that cells undergo when

shed off their central mass, leads to eventual implantation and metastasis. The process of

evading apoptosis is integral to the assist the development of secondary, metastatic tumours

and is important in the removing the reliance of GJIC to maintain cell integrity. Researchers

became aware that apoptosis, a form of cell death, played a crucial role in a myriad of

physiological and pathological processes (Kerr et al. 1972) Apoptosis is often referred to as a

physiological cell suicide program that is critical for the development and maintenance of

healthy tissues (Deveraux et al. 1999). The mechanisms for apoptosis have been strongly

conserved during evolution and dysregulation of cell death pathways occur in cancer,

Page 51: Investigating the effects of, the isoflavone, Phenoxodiol

31

autoimmune and immunodeficiency diseases and in neurodegenerative (Deveraux et al. 1999).

The process of apoptosis is necessary to remove unwanted cells from multicellular organisms.

It is the mechanism by which specific groups of cells are removed during embryogenesis so

that a particular course of development may be followed. The commitment to apoptosis

involves both signalling and effector stages. Though they vary between cells, the array of

apoptosis induction signals trigger signalling pathways that coalesce, often at the

mitochondria, to activate central effector pathways involving a series of proenzymes, the

caspases (Maguire et al. 2000). When activated caspases are efficient at degrading cellular

processes, and DNA, to the point where normal physiological activity is impossible, resulting in

cellular response and apoptotic phenomena (Hengartner 2000).

2.5.1. MORPHOLOGY OF APOPTOTIC CELLS

Apoptosis occurring in cells is characterised by common distinctive morphological and

molecular features (Kerr et al. 1972). Apoptotic cells will undergo a regulated autodigestion,

which involves the disruption of cytoskeletal integrity, cell shrinkage, nuclear condensation,

activation of endonucleases, blebbing of the cell surface, chromatin condensation and the

formation of apoptotic bodies distinct to apoptosis (Maguire et al. 2000; Birkey Reffey et al.

2001). The result is contraction of the cytoplasm, accompanied by condensation of nuclear

chromatin into several large masses. Caspase mediated events are likely to contribute to the

characteristic morphological changes that result in DNA cleavage at the internucleosomal

region into 185 base pair multiples (Hengartner 2000). After fragmentation, DNA exhibits a

typical condensation at the nuclear margin. The cell forms membrane bound apoptotic bodies

of various sizes, containing organelles and nuclear fragments. These apoptotic bodies provide

a strong stimulus for phagocytosis and are subsequently consumed by their viable

neighbouring cells or specialist phagocytes (Schwartzman and Cidlowski 1993).

Page 52: Investigating the effects of, the isoflavone, Phenoxodiol

32

The process of apoptosis is a highly conserved mechanism in multicellular organisms which

allows for the removal of unwanted cells without an inflammatory response (Deveraux et al.

1999; Hengartner 2000). Once the apoptotic process is triggered cells undergo a series of

molecular and morphologic changes that characterise classical apoptosis. These processes

include; irreversible chromosome condensation (pyknosis 1 ) with nuclear fragmentation

(karyorrhexis2) via inter-nucleosome DNA cleavage, cell shrinkage, formation of multiple

membrane blebs (zeiosis3) and finally the breakdown of the cell into apoptotic bodies

containing organelles and pieces of the degraded nucleus that are phagocytosed by

surrounding cells (van Heerde et al. 2000). Recently, necrosis, once thought of as simply a

passive, unorganised way to die, has emerged as an alternate form of programmed cell death

whose activation might have important biological consequences, including the induction of an

inflammatory response. Autophagy has also been suggested as a possible mechanism for non-

apoptotic death despite evidence from many species that autophagy represents a survival

strategy in times of stress. Recent advances have helped to define the function of and

mechanism for programmed necrosis and the role of autophagy in cell survival and suicide

(Edinger and Thompson 2004).

Other apoptotic modifications include early membrane changes such as externalisation of

membrane bound phosphatidylserine which promotes the phagocytosis of the apoptotic

bodies without inducing an inflammatory response (Martin et al. 1995). This early

externalisation is a key step in detecting cells entering early stage apoptosis and several

distinct apoptotic events have been identified on the molecular signalling level following

phosphatidylserine externalisation including; the signalling stimulus for degradation of DNA

1 Pyknosis or karyopyknosis, is the irreversible condensation of chromatin in the nucleus of a cell

undergoing necrosis or apoptosis. It is followed by karyorrhexis 2 Karyorrhexis The destructive fragmentation of the nucleus of a dying cell whereby its chromatin is

distributed irregularly throughout the cytoplasm 3 Zeiosis is the term used to describe the formation of blebs in a cell. A bleb is an irregular bulge in the

plasma membrane of a cell caused by localised decoupling of the cytoskeleton from the plasma membrane.

Page 53: Investigating the effects of, the isoflavone, Phenoxodiol

33

into fragments of multiples of ~200 base pairs, the proteolytic cleavage of poly(ADP-ribose)

polymerase (PARP) and cytoskeleton components separation (van Heerde et al. 2000).

2.5.2. APOPTOTIC SIGNALLING

Apoptosis is regulated at many levels, including the initiation, transduction, amplification and

execution stages and the mutations that disrupt each stage have been detected in tumour

cells. Because mutations in cancers necessarily produce a selective advantage to emerging

tumour cells, the identification of mutated components and their frequency of mutation

highlight critical regulatory points in survival and proliferative processes. The fact that

apoptosis is disabled at distant stages in different tumour types suggest that its critical control

points are context dependent. Identification of these control points singles out distinct “sites

of attack” for targeting by novel chemotherapeutic drugs (Johnstone et al. 2002).

On the molecular level, the cell death program can be divided into three parts: initiation,

execution, and termination of apoptosis. Apoptosis is initiated by a variety of stimuli, including

growth factor withdrawal (“death by neglect”), UV or -irradiation, chemotherapeutic drugs,

and death receptor signals. In most cases the execution phase is characterised by membrane

inversion and exposure of phosphatidylserine, blebbing (zeiosis), fragmentation of the nucleus,

chromatin condensation, and DNA degradation. In the termination phase, “apoptotic bodies”

are engulfed by phagocytes resulting in no inflammatory response (Hengartner 2000).

While mutations in cancer cells often target regulators of the intrinsic apoptotic pathway

(indirect) such as p53 and the Bcl-2 related proteins, alterations that disrupt apoptosis

downstream of the mitochondria have been reported (Petronilli et al. 2001). Silencing of Apaf-

1 occurs in metastatic melanoma and over expression of IAPs and heat shock proteins (HSP),

which can inhibit Caspase-9 activation, is commonly observed in human tumours (Deveraux et

al. 1999). Postmitochondrial mutations appear less frequently than those targeting upstream

Page 54: Investigating the effects of, the isoflavone, Phenoxodiol

34

components of the apoptotic program, though this could represent greater redundancy in the

downstream pathway or difficulty maintaining cell viability following damage to the

mitochondria. Consistent with this, apoptosis can be induced without the activation of

caspases or while the caspases themselves are prevented from action by a pan caspase

inhibitor such as Z-VAD-FMK (Johnstone et al. 2002).

2.5.3. EXTRINSIC PATHWAYS

To date, it is accepted that every cell type harbors the machinery to commit suicide by

apoptosis. Subsequently, it is acknowledged that apoptosis plays a crucial role in homeostasis

and pathology and the molecular biology and biochemistry of apoptotic death machinery are

far from being completely resolved (van Heerde et al. 2000). Despite apoptosis exhibiting great

diversity in signalling pathways, three functionally distinct phases of apoptosis, common to all

cell types can be distinguished. First, the initiation phase can be induced by death inducing

signals such as Fas ligand (Fas-L) and tumour necrosis factor α (TNF-α), a lack of growth and/or

survival signals, or DNA damage which may devise the cell to prepare for suicide, commonly

these pathways involve the Caspase signalling system (Susin et al. 2000; Cregan et al. 2004;

Dohi et al. 2004; Svingen et al. 2004). The initiation phase will result in the activation of the

second more general, decision phase, in which the cell is still able to make the decision to live.

This phase is characterised, in most cases, by involvement of the mitochondrion. This organelle

provides molecular links between the upstream initiation phase and downstream execution

phase, by releasing apoptosis inducing factor (AIF), Cytochrome c, Caspases-2,-3 and -9. When

the cell is committed to die, and thus the point of no return has been passed, the third phase

the execution phase is activated (Wang et al. 1999). This phase is characterised by the

activation of the downstream or effector caspases which subsequently orchestrate a sequence

of events by their hierarchical activation. These events include the loss of cell junctions, cell

Page 55: Investigating the effects of, the isoflavone, Phenoxodiol

35

shrinkage, chromatin condensation and margination, nuclear pyknosis and fragmentation,

membrane blebbing and disassembly of the cell into membrane-enclosed vesicles called

apoptotic bodies. Several events have been identified on the biochemical level, including the

degradation of DNA into fragments of multiples of ~200 base pairs, the proteolytic cleavage of

poly(ADP-ribose) polymerase (PARP) and cytoskeleton components, and the cell surface

exposure of phosphatidylserine (PS) (van Heerde et al. 2000).

2.5.4. INTRINSIC PATHWAYS

Disruption of the intrinsic pathway is common in cancer cells with the p53 tumour suppressor

gene the most frequently mutated gene in human tumours, and loss of p53 function can both

disable apoptosis and accelerate tumour development in transgenic mice (Ryan et al. 2001).

Function mutations or altered expression of p53 downstream effectors (PTEN, Bax, Bak, and

Apaf-1), or upstream regulators (ATM, Chk2, Mdm2, and p19ARF), occur in human tumours

and as a result, the presence of wild-type p53 does not necessarily indicate that the pathway is

intact and therefore correlating p53 gene integrity with a functional p53 pathway is not always

correct (Schmitt et al. 1998). The B-cell lymphoma 2 (Bcl-2) family members are key regulators

of the intrinsic apoptotic pathway and consist of both pro- and anti-apoptosis members who

expression is commonly altered in tumours (Reed 2000). Mutations or altered expression of

upstream regulators of Bcl-2 proteins are associated with cancer. For example, the Bad-kinase

intrinsic Akt, is positively regulated by various oncoproteins, and negatively regulated by the

PTEN tumour suppressor, amplified Akt and mutated PTEN, have been found with high

frequency in a variety of solid cancers, indicating the importance of intrinsic cell signalling in

regulating tumourigenisis (Datta and Datta 1999).

Page 56: Investigating the effects of, the isoflavone, Phenoxodiol

36

2.5.5. CASPASES

Caspases (Cysteine Aspartases) are the key effector proteases of apoptosis, existing in healthy

cells as inactive precursor molecules (zymogens) called procaspases (Silke et al. 2001;

Donepudi and Grutter 2002). Three major apoptotic pathways have been identified: one

activated by death receptor activation and the other by intrinsic stress induced stimuli and

finally a endoplasmic reticulum specific stress pathway has been determined (Nakagawa et al.

2000; MacFarlane et al. 2002). Mechanisms described for activating caspases include;

noncovalent association with caspase activating proteins, such as Apaf-1 or FAS-L, leading to

autocatalytic cleavage of the procaspase polypeptide at specific aspartic acid residues, or

cleavage at specific aspartic acid residues within the inactive zymogens by other activated

caspases (Zou et al. 2003). The caspases are highly conserved throughout the animal kingdom

(Deveraux et al. 1997), and many function to initiate Cytochrome c release from the

mitochondria.

When added to cytosol environment Cytochrome c initiates an apoptotic program which

involves proteolytic processing and activation of more caspases, and apoptotic-like destruction

of exogenously added nuclei (Deveraux et al. 1997). Caspase functions are subject to

modulation by a family of inhibitor of apoptosis proteins (IAP’s) (Deveraux et al. 1999). Effector

caspases, such as Caspase-3 and Caspase-7 are activated by initiator caspases, such as

Caspase-9 and Caspase-8, through proteolytic cleavage. Once activated, the effector caspases

are responsible for the proteolytic cleavage of a diverse array of structural and regulatory

proteins, resulting in the apoptotic phenotype (Suzuki et al. 2001). The initiator caspases

Caspase-9 and Caspase-8 converge on the activation of the executioner caspases, Caspase-3

and Caspase-7. The critical role of those caspases in transmitting cell death signals is

underscored by the results of many phenotypic gene ablation experiments that vary between

embryonic lethality (Caspases-7 and -8) to perinatal lethality (Caspases-3 and -9) (Riedl and

Page 57: Investigating the effects of, the isoflavone, Phenoxodiol

37

Salvesen 2007). Caspases-3 and -7 are considered to be the major effector caspases within

cells, activated by the initiator caspase cascade, they can feed back to alter upstream caspases

such as Caspase-8 and Caspase-9 allowing some degree of self regulation (Silke et al. 2001).

FIGURE 9 SIMPLE APOPTOTIC SIGNALLING ADAPTED FROM RIEDL AND SALVESEN 2007

The pathways in evidence are the caspase-dependant apoptotic signalling resulting from

extrinsic (DISC) or intrinsic (apoptosome) stimuli. The proteolytic caspase cascade in the

initiation and execution of apoptosis. Both the intrinsic and extrinsic pathways use related

principles in sensing an apoptotic signal and executing apoptosis. In the intrinsic pathway, an

apoptotic stimulus in the cell leads to the assembly of a signalling platform, the apoptosome,

which activates the initiator Caspase-9. An extrinsic apoptotic signal, by contrast, is mediated

by binding of an extracellular ligand to a transmembrane receptor, leading to the formation of

the death-inducing signalling complex (DISC), which is capable of activating the initiator

Caspase-8. Once activated, either Caspase-9 or Caspase-8 cleaves executioner Caspase-3 and

Caspase-7, which represents the execution level of the caspase cascade and leads to apoptosis

of the doomed cell. Negative regulators of the caspase cascade can be found on both levels.

Page 58: Investigating the effects of, the isoflavone, Phenoxodiol

38

Whereas FLIP blocks the activation of the initiator Caspase-8 in the DISC, XIAP can block both

the initiation phase, by inhibiting Caspase-9, and the execution phase, by blocking Caspase-3

and Caspase-7, of the cascade. Adapted from (Riedl and Salvesen 2007).

2.5.6. APOPTOSOME

Upon intrinsic stimulation of apoptosis, the release of Bax and Cytochrome c from the

mitochondria initiates the apoptosis promoting activity of Cytochrome c by allowing

interaction with the Apaf-1 protein. Binding of Cytochrome c to Apaf-1 causes recruitment of

Caspase-9, forming the apoptosome (Slee et al. 1999; Slee et al. 1999; Srinivasula et al. 2001).

The apoptosome is a large molecule that interacts with the executioner caspases, Caspase-3

and Caspase-7, which initiate DNA fragmentation and cause the cell to enter into apoptosis

(Boatright and Salvesen 2003). Caspase-3 can be activated via extrinsic pathways that produce

activated Caspase-9 through Caspase-8 signalling. If intrinsic pathways are activated, the

release of Cytochrome c and apoptosome formation results in Caspase-3 activation (Riedl and

Salvesen 2007).

2.5.7. BCL-2 FAMILY

The mitochondrial pathway of cell death is mediated by Bcl-2 family proteins, a group of anti-

apoptotic and pro-apoptotic proteins that regulate the passage of small molecules, such as

Cytochrome c, Smac/Diablo, and Apoptosis Inducing Factor (AIF), which activates caspase

cascades, through the mitochondrial transition pore. The activation of caspases is

counteracted by anti-apoptotic molecules of the Bcl-2 family (Bcl-2, Bcl-xL), because these Bcl-

2 family proteins heterodimerise with proapoptotic members of the Bcl-2 family (Bax, Bak) and

interfere with release of cytochrome c by pore-forming proteins (Bid, Bik) (Gross et al. 1999).

Page 59: Investigating the effects of, the isoflavone, Phenoxodiol

39

B-cell lymphoma-extra large (Bcl-xL) is a transmembrane molecule in the mitochondria. It is

involved in the signal transduction pathway of the FAS-L and is one of several anti-apoptotic

proteins which are members of the Bcl-2 family of proteins. It has been implicated in the

survival of cancer cells with Bcl-xL expression in prostate cancer leading to cell proliferation,

apoptosis inhibition and obvious decreased radiosensitivity in human prostatic carcinoma cells

(Wang et al. 2008). In the absence of activated Akt, Bcl-2 family member Bad forms

heterodimers with Bcl-xL, an antiapoptotic protein that prevents the release of Cytochrome c

from mitochondria (Gross et al. 1999). This complex formation abrogates the antiapoptotic

function of Bcl-xL thereby facilitating apoptotic death via a cytochrome c-dependent pathway.

As a consequence, the dynamic interaction between Bcl-xL and Bad represents a critical

determinant of cell fate downstream of the phosphatidylinositol-3 kinase (PI3K)/Akt cascade,

and may represent an alternative mechanism for cancer cells to evade apoptosis (Cheng et al.

2001; Lee et al. 2008).

Bax is an important multidomain pro-apoptotic proteins essential for the mitochondrial release

of Cytochrome c. The expression of Bax is up regulated by the tumour suppressor protein p53,

and Bax has been shown to be involved in p53-mediated apoptosis (Goodman et al. 1998). In

normal mammalian cells, the majority of Bax is found in the cytosol, but upon initiation of

apoptotic signalling, Bax undergoes a conformational shift, and inserts into organelle

membranes, primarily the outer mitochondrial membrane resulting in Cytochrome c release

(Lee et al. 2008).

2.5.8. X-LINKED INHIBITOR OF APOPTOSIS PROTEIN

X-linked inhibitor of apoptosis protein (xIAP) is part of a family of endogenous caspase

inhibitors that, when up regulated, have been implicated in the early stage development of

Page 60: Investigating the effects of, the isoflavone, Phenoxodiol

40

prostate and breast epithelial carcinoma (Parton et al. 2002; Krajewska et al. 2003). In

conjunction with anti-apoptotic Bcl-2 members (Bcl-2, Bcl-xL) IAP family members exert

significant anti-apoptotic effects through; inhibition of Caspase-3 and -7 activation and

function, inactivation of pro-apoptotic Bcl-2 family members and involvement in signalling

pathways NF-κB and JNK (Levkau et al. 2001; Sanna et al. 2002; Hu et al. 2003; Dubrez-Daloz et

al. 2008). XIAP expression levels are also found to be increased in epithelial ovarian and

prostate cancers, and have been linked to chemoresistance and progression of cells to

metastasis (Krajewska et al. 2003; Berezovskaya et al. 2005; Mahoney 2007).

As opposed to other members of the IAP family, XIAP is the only member capable of direct

inhibition of both the execution and initiation phases of the apoptotic cascade; it inhibits

apoptotic activation by inhibiting caspase-9, and inhibits the execution phase by inhibiting the

effector caspases, Caspase-3 and Caspase-7 (Kluger et al. 2007). As XIAP selectively binds and

inhibits Caspase-3, -7 and inhibits Apaf-1-Cytochrome c-mediated activation of Caspase-9 (the

apoptosome) it is responsible for directly inhibiting much of the apoptotic pathways and is a

considered a key factor in chemoresistance in tumours (Mahoney 2007).

2.5.9. APOPTOSIS INDUCING FACTOR

Apoptosis Inducing Factor (AIF) is one of the apoptogenic molecules released from the

mitochondria upon multiple stimuli with overexpression of AIF causing apoptosis independent

of caspases (Xie et al. 2005). AIF has also been shown to act as a free radical scavenger and

play a role in normal mitochondrial oxidative phosphorylation. Upon stimulus by some death

signals AIF translocates to the nucleus when apoptosis is induced (Susin et al. 1999).

Experiments using recombinant AIF showed that it induced chromatin condensation in isolated

nuclei and large-scale fragmentation of DNA, it induced purified mitochondria to release the

Page 61: Investigating the effects of, the isoflavone, Phenoxodiol

41

apoptogenic proteins Cytochrome c and Caspase-9 and microinjection of AIF into the

cytoplasm of intact cells induces condensation of chromatin, dissipation of the mitochondrial

transmembrane potential, and exposure of phosphatidylserine in the plasma. None of these

effects is prevented by the wide-ranging caspase inhibitor known as Z-VAD-FMK indicating that

AIF is a mitochondrial effector of apoptotic cell death independent of caspase activity (Susin et

al. 1999).

2.5.10. REACTIVE OXYGEN SPECIES

Chemotherapy and radiation treatments are known to induce oxidative stress in cancer cells,

rapid accumulation of highly reactive molecules, such as, nitric oxide occurs and results in

damage to cell structures and potentially activates the apoptotic cascade. Oxidative stress

induces the expression of antioxidant genes that contain an antioxidant response element in

their promoters; this can result in a triggering of apoptosis or necrosis of various cell types and,

in several instances, can inhibit cell growth and interfere with the cell cycle (Minelli et al.

2009). Cancer cells protect themselves from reactive oxygen species (ROS) by increasing the

amount of proteins like superoxide dismutase (SOD) which quickly responds to ROS production

and removes it (Murphy 2009). In this regard, it has been reported that Bcl-2 reduces

accumulation of reactive oxygen species (ROS) in transfected cells and can protect a variety of

cells from apoptosis induced by oxidative stressors. Conversely, the Bcl-2-related protein Bax

causes increased mitochondrial production of ROS and apoptosis when overexpressed in

cultured cells (Dharmarajan et al. 1999). Aside from Bcl-2, other factors that play a role in

protecting cells from oxidative stress include superoxide dismutases (SOD), which are

responsible for the conversion of superoxide radical anion to peroxide intermediates, and

glutathione peroxidase and catalase, which convert peroxides to water (Dharmarajan et al.

1999).

Page 62: Investigating the effects of, the isoflavone, Phenoxodiol

42

2.6. THE CELL CYCLE

The cell cycle is composed of G1, S, G2 and M phases, which represent normal function, DNA

replication, organelle replication and mitotic separation respectively (Marieb 2012). Any errors

that may occur at these steps could be catastrophic for normal cell functioning and, as such,

the cell cycle apparatus retains potent signalling molecules that can search for errors and

rapidly induce apoptosis (Cooper 2003). The initial point where this process can occur is

referred to as the G1 restriction point (Pardee 1974).

To maintain normal function, the cell cycle is a potent inducer of cell death through regulation

of DNA content at a variety of restriction points. As the cell reaches a restriction point, the

gene p53 detects and attempts to repair any evident DNA damage; if the damage is too great it

can initiate apoptosis. The p53 gene is a potent tumour suppressor that exerts its functions

through activation of downstream targets, some of which include induction of CDKN1/p21WAF1,

14-3-3y, and REPRIMO for G1-G2 arrest; p53R2 for DNA repair and Bax, Puma, p53AIP1, PERP,

and CD95 for apoptosis. Unfortunately p53 is found mutated in nearly 70% of carcinomas(el-

Deiry et al. 1993; Scherr et al. 1999; Ryan et al. 2001; Fojo and Bates 2003; Huo et al. 2004).

Considerable evidence indicates that the choice made by p53 to activate cell cycle arrest and

DNA repair pathways, or the apoptosis pathway after DNA damage, is dependent on the

extent of unrepaired or misrepaired double-strand breaks in the DNA (Devlin et al. 2008).

Disabling the p53 pathway enables cells to enter and proceed through the cell cycle under

conditions that increase the frequencies of aneuploidy, gene amplification, deletion and

translocation. This, coupled with loss of p53 dependant apoptosis, increases genetic instability

and is highly selected during cancer progression (Vafa et al. 2002).

Page 63: Investigating the effects of, the isoflavone, Phenoxodiol

43

2.6.1. CYCLINS, CYCLIN DEPENDANT KINASES

Cell cycle progression is intricately regulated by the interactions of cyclin and cyclin-dependent

kinase (Cdk) complexes. Cyclin-D1 is a regulatory subunit of the highly conserved cyclin family

that phosphorylates and, together with sequential phosphorylation by cyclin-E/CDK2,

inactivates the cell-cycle inhibiting function of the retinoblastoma protein. Upon mitogenic

stimulation, the cyclin-D Cdk4/Cdk6 and cyclin-E Cdk2 complexes mediate phosphorylation of

retinoblastoma protein, which induces transcriptionally active E2F thereby ensuring G1-S

transit (Baldin et al. 1993; Roy et al. 2007). Retinoblastoma protein serves as a gatekeeper of

the G1 phase, and passage through the restriction point leads to DNA synthesis, making cyclin-

D1 expression an important target to regulate cancer cell replication (Fu et al. 2004).

2.6.2. P21WAF1/P53

p21WAF1 is a cyclin dependant kinase inhibitor (CdkI) family member, along with p27 and p57,

which interfere with the cyclin dependant kinase-cyclin complex. The group is regulated both

by internal and external signals, with the expression of p21WAF1 under transcriptional control of

the p53 tumour suppressor gene (Srivastava et al. 2007). As an important mediator of

apoptosis, the frequency of somatic mutations in the p21WAF1 gene and family in cancers is very

rare; which underlines the importance of these molecules as promising therapeutic targets.

Both p21 and p27 are upregulated by a variety of regulatory pathways at the transcriptional as

well as post-transcriptional levels, with p21WAF1 transcriptionally upregulated by p53 in

response to DNA damage. p21WAF1 is also activated by various transcription factors and

subsequently mediates growth arrest, senescence and apoptosis in a p53-independent

manner, and p21 mRNA stability can be post transcriptionally regulated by HuR, an RNA-

binding protein, in response to stress (Roy et al. 2007; Roy et al. 2008).

Page 64: Investigating the effects of, the isoflavone, Phenoxodiol

44

p21WAF1 is a potent cyclin-dependent kinase inhibitor (CKI). The p21 (CIP1/WAF1) protein binds

to and inhibits the activity of cyclin-CDK2 or -CDK1 complexes, and thus functions as a

regulator of cell cycle progression at G1. The expression of this gene is tightly controlled by the

tumour suppressor protein p53, through which this protein mediates the p53-dependent cell

cycle G1 phase arrest in response to a variety of stress stimuli. This was a major discovery in

the early 1990s that revealed how cells stop dividing after being exposed to damaging agents

such as radiation. In addition to growth arrest, p21 can mediate cellular senescence and one of

the ways it was discovered was as a senescent cell-derived inhibitor (Niculescu et al. 1998;

Vafa et al. 2002). Expression of p21 is mainly dependent on two factors 1) stimulus provided

2) type of the cell. Growth arrest by p21 can promote cellular differentiation. p21 therefore

prevents cell proliferation. p53 exerts its functions mainly through transactivational activity,

including the induction of CDKN1/p21, 14-3-3y, and REPRIMO for G1-G2 arrest; p53R2 for DNA

repair and Bax, Puma, p53AIP1, PERP, and CD95 for apoptosis. Considerable evidence indicates

that the choice made by p53 to activate cell cycle arrest and DNA repair pathways or the

apoptosis pathway after DNA damage is dependent on the extent of unrepaired or misrepaired

double-strand breaks in the DNA (Waldman et al. 1996; Devlin et al. 2008).

2.6.3. KI-67

Ki-67 antigen is present in all proliferating cells (normal and neoplastic) and its evaluation

allows determining growth fraction of cellular population in a relatively easy way. Previous

studies proved the hypothesis that in breast cancer the Ki-67 index is an independent

prognostic factor in both patient survival assessment and disease recurrence. Ki-67 index can

also be used as a predictive factor of neoplastic cell response to certain types of therapy (Koda

et al. 2007). An important indicator of cell cycle pace, Ki-67 is present in all proliferating cells,

normal and neoplastic, and its level of expression indicates the rate of growth of a cell

population. Studies proved the hypothesis that, in breast cancer, the Ki-67 index is an

Page 65: Investigating the effects of, the isoflavone, Phenoxodiol

45

independent prognostic factor in both patient survival assessment and disease recurrence

(Dudderidge et al. 2007).

2.6.4. C-MYC

c-Myc is an important regulator of cell function that was implicated as one of the first

oncogenes. It is known that cell cycle regulation is altered under excess c-Myc expression with

a decrease in time taken to reach the restriction point of G1 (Yin et al. 1999). c-Myc protein

overexpression has been reported to immortalise cells, to reduce their growth factor

requirements and to promote cell cycle progression and genomic instability (Wasylishen and

Penn 2010). According to a role in the maintenance of the mitotic spindle integrity,

overexpression of c-Myc has been reported to disrupt the spindle checkpoints activated by

taxanes. In colon cancer cell lines, c-Myc amplification has been related to the modulation of

the multiple effects of paclitaxel (Cassinelli et al. 2004). C-Myc cell cycle regulation has been

implicated by varied means; canonical Wnt/β-catenin stimulation, Cyclin-D and Cyclin-E CDK2

stimulation, inhibiting the effect of p27 a CDKI family member amongst others (Mateyak et al.

1999).

2.7. WNT SIGNALLING

The Wnt family of secreted ligands operates through multiple receptors, to modulate discrete

intracellular linear and integrated signalling pathways in embryonic development, in adults and

in prostate cancer progression. Canonical signalling occurs when a Wnt ligand attaches to a

Frizzled (Fzd) family receptor and to a low density lipoprotein receptor-related protein 5 (LRP5)

or LRP6 coreceptor. This triggers a cytoplasmic Wnt/β-catenin transduction cascade, which

Page 66: Investigating the effects of, the isoflavone, Phenoxodiol

46

mediates β-catenin stability via glycogen synthase kinase 3 beta (GSK3β) and results in a

context dependent transcriptional outcome. Stabilized β-catenin forms multicomponent

transcriptional complex with LEF⁄Tcf and activates downstream targets such as c-Myc (Farooqi

et al. 2011). In the absence of Wnt signals, cytoplasmic β-catenin is sequestered by a complex

consisting of Axin, the tumour suppressor Adenomatous polyposis coli (APC), and GSK3β (Hall

et al. 2006).

Non-canonical Wnt proteins such as Wnt5a, lead to the activation of protein kinase C and

calcineurin with downstream intracellular targets. The Wnt/Calcium pathway has been

reported to control proliferation and have the ability to act as a tumour suppressor (Liang et al.

2003). The activity of Wnt proteins is controlled by soluble extracellular antagonists including

secreted frizzled-related proteins (sFRP), Wnt inhibitory factor-1 (Wif1), Cerberus, and

Dickkopf (DKK) (Hirata et al. 2011). sFRP, WIF-1, and Cerberus act as competitive inhibitors of

the frizzled receptor by sequestering Wnt factors and can, therefore, block both canonical and

non-canonical Wnt pathways. DKK-1, in contrast, binds the Wnt co-receptors LRP 5 and 6 to

block canonical Wnt signalling (Hall et al. 2010).

2.7.1. Β-CATENIN

The protein β-catenin has at least 2 functions of interest in prostate cancer: it participates in

cadherin-mediated adhesion, and it is the "molecular node" of the Wnt canonical signalling

pathway. In the absence of Wnt signals (i.e., when the pathway is inactive), the

serine/threonine kinase glycogen synthase kinase 3 beta (GSK3β) forms complexes with

proteins, which in turn bind to soluble β-catenin and facilitate its phosphorylation (Wan et al.

2012). Phosphorylated β-catenin then binds an E3 ubiquitin ligase then undergoes

proteosomal degradation, preventing the accumulation and transcriptional activity of β-

Page 67: Investigating the effects of, the isoflavone, Phenoxodiol

47

catenin. When Wnt ligands bind their receptor complex, the resulting activation of the

cytoplasmic protein dishevelled inactivates GSK3β, thereby preventing degradation of soluble

β-catenin and stabilizing it in the cytoplasm (Hirata et al. 2011; Menezes et al. 2012).

Cytoplasmic β-catenin then translocates to the nucleus, where it heterodimerises with

transcription factors of the T-cell factor/lymphoid enhancer–binding factor (TCF/LEF) family

(van Es et al. 2003). Accumulation of soluble β-catenin is therefore critical for activation of Wnt

transcription in the pathway. More recently, another group reported that activation of Wnt/β-

catenin signalling is involved in prostate cancer initiation and progression in a mouse model

(Yu et al. 2011). Together, these findings imply that the Wnt canonical pathway is involved in

the pathogenesis of a subgroup of advanced prostate cancers.

2.7.2. ANDROGEN RECEPTOR AND Β-CATENIN

Initiating the normal and cancerous prostate cell cycle progression is the activity of the growth

factor androgen receptor (AR), which is found to be mutated in most late stage prostate

cancers. The Wnt pathway and its interaction with AR have been suspected to play important

roles in prostate cancer (Wang et al. 2008). Initiating the cell cycle progression is the activity of

the growth factor androgen receptor (AR), which is found to be mutated in most late stage

prostate cancers and is widely accepted to be the initiator of prostate cell replication and

differentiation through transcription of mitogenic and anti-apoptotic. Pathways such as the

canonical Wnt/β-catenin signalling pathway drive the cells towards growth and differentiation

while interaction with the anti-apoptotic Bcl-xL prevent the cells from entering into apoptosis

when AR is stimulated (Sun et al. 2008; Wang et al. 2008). Because AR signalling has also been

shown to play a key role in prostate carcinogenesis, androgen ablation therapy is a commonly

used form of treatment, particularly for advanced disease. While ablation therapy leads to

initially significant levels of prostate cancer cell apoptosis, the effect is short-lived and

Page 68: Investigating the effects of, the isoflavone, Phenoxodiol

48

ultimately not curative as most patients develop androgen-independent disease (D'Antonio et

al. 2008). Manipulation of this growth factor by; upregulation of receptors, mutation to a

constitutively active form or a complete loss of AR as a mitogenic requirement are

characteristics of late stage prostate cancer (Peng et al. 2008).

Levels of β-catenin and AR are both increased in hormone refractory prostate cancer, β-

catenin is activated by canonical Wnt stimulation, but mutated forms of β-catenin, which can

result in a permanently stabilised protein, have also been detected in prostate cancer. β-

catenin has multiple functions that involve both cell adhesion and signal transduction in

response to Wnt ligands (Chesire et al. 2002). In the absence of Wnt ligands, GSK3β complexes

with other proteins and degrades cytoplasmic β-catenin, but when Wnt ligands bind to the

frizzled receptor complex, the resulting activation of the cytoplasmic protein dishevelled

inactivates GSK3β, thereby preventing degradation of β-catenin (Wan et al. 2010; Wan et al.

2012). It is important to note that GSK3β has been linked to prevention of AR transcriptional

activity; so Wnt ligand binding both increases AR expression and β-catenin stability (Li et al.

2008). Upon stimulation, β-catenin translocates to the nucleus where it can drive cell cycle

progression through interaction with T-cell factor (TCF)/lymphoid enhancer factor (LEF)

transcription factors to initiate transcription of target genes such as c-Myc and Cyclin-D1,

driving cell growth. As such, dysregulation of the Wnt, AR or β-catenin pathways can result in a

metastatic, highly replicative phenotype (Wang et al. 2008). It has been postulated that the

ratio of β-catenin/AR might be an important prognostic indicator that may even help define a

subpopulation of men with prostate cancer for individualised management (Wan et al. 2012).

2.7.3. SFRP4

Page 69: Investigating the effects of, the isoflavone, Phenoxodiol

49

Secreted frizzled related protein 4 (sFRP4) is one of a group of proteins that antagonise the

canonical Wnt/β-catenin pathway, decreasing Wnt activity and preventing activated β-catenin

from forming (Drake et al. 2003; Drake et al. 2009). sFRP4 can also decrease invasiveness in

androgen-independent prostate cancer cells and has been shown to be anti-angiogenic and

pro-apoptotic in nature (Muley et al. 2010). Moreover, the correlation between increased

membranous sFRP4 and β-catenin expression in a large human cohort supports evidence for

sFRP4 as a prognostic marker in localised androgen-dependent prostate cancer (Horvath et al.

2004). Unlike other inhibitors of Wnt signalling, sFRP4 appears to affect androgen-dependent

and androgen-independent prostate cancer (Horvath et al. 2007).

Page 70: Investigating the effects of, the isoflavone, Phenoxodiol

50

3 . C H A P TER T H REE: MA T ERIA L S A N D METH OD S

3.1. TISSUE CULTURE

Cell culture was performed in a tissue culture facility located within a physical containment

level 2 (PC2) grade laboratory in the school of Anatomy, Human Biology and Physiology at the

University of Western Australia. The tissue culture facility utilised biohazard class II hoods and

cell culture incubators operating at 37°C 5% CO2. Incubators were sterilised six monthly to

prevent contamination and water in the 37°C water bath and incubators was treated with anti-

bacterial agent Aqua-clear (Cleanware).

3.2. CELL CULTURE

All cell lines were cultured in T-75cm2 tissue culture flasks with vented, filtered caps. Two types

were used over the length of the project; one from Sarstedt (cat#83.1813.002) and one from

Corning (cat#3290). Similarly 96 well plates (Sarstedt cat#83.1835.500, Corning cat#3595) and

6 well plates (Sarstedt cat#83.1839.500, Corning cat#3516) were bought from both companies

with no detectable change in cellular adherence to plastic or growth rate. Tissue culture media

was replaced every three days with cells reaching confluence approximately every five to eight

days dependent upon seed rate. Flasks were reused for 4 passages before being replaced by a

new flask. 10cm dishes were supplied by BD Falcon (cat#353003).

Page 71: Investigating the effects of, the isoflavone, Phenoxodiol

51

3.2.1. SUBCULTURE AND COUNTING OF CELLS

All cell lines were subcultured in a similar manner. Cells had their media aspirated, 4ml of 1x

Phosphate Buffer Solution (PBS) was added to the flask then aspirated off. Following the wash

with 1xPBS, 3ml Trypsin 0.5% EDTA (Gibco Cat#25300) was quickly washed over the cells

before being aspirated and another 3ml of Trypsin 0.5% EDTA was added, cells were moved to

a 37°C 5% CO2 incubator for 5 minutes or until the cells had released off the surface. Flasks

were given a gentle tap to encourage cell separation and to assist the release of adherent cells

from the plastic surface. The trypsin wash step noticeably increased the speed at which cells

would successfully release from the flask plastic.

Following trypsinisation, 5ml of complete growth media was added into the flask, to inactivate

the trypsin EDTA mixture, and the mixture was washed up and down along the flask’s base

surface, increasing cell homogenisation and washing off the last adherent cells, before being

transferred to a 15ml Falcon tube (Sarstedt Cat#62.554.205-500CS). The cells were mixed

thoroughly to ensure homogeneity then 20µL was removed and placed into both sides of a

haemocytometer. The inside 5x5 well section of the haemocytometer was counted on both

sides, with the top and right line touching cells counted while the bottom and left touching

cells were discarded. An average of the two numbers was taken and then multiplied x104 to

get the concentration of cells/mL in suspension.

3.2.2. CELL CRYOPRESERVATION AND STORAGE

To ensure adequate numbers of low passage cells, stocks were frozen down regularly for long

term use. Cells were trypsinised and counted as previously described, and then the 15ml

Falcon tube was centrifuged at 1.5krpm/0.4rcf on an Eppendorf 5702 centrifuge for 3 minutes.

Page 72: Investigating the effects of, the isoflavone, Phenoxodiol

52

Following centrifugation the cells were clumped in the bottom of the tube and the

media/trypsin aspirated off. Freezing down solution was added such that the cells were left in

a concentration of 1x106 cell/mL. Freezing down solution was a combination of 90% total

growth media and 10% dimethyl sulfoxide (DMSO, Sigma cat#D2650). Cells were placed in a

freezing down container (Nunc) containing -20°C isoproanol and moved into a -80°C freezer for

24 hours. After being in the -80°C freezer for 24 hours the cells were removed and placed in a

liquid nitrogen long term storage container.

3.2.3. LNCAP CELLS

The LNCaP clone was isolated in 1977 by J.S. Horoszewicz, from a needle aspiration biopsy of

the left supraclavicular lymph node of a 50 year old Caucasian male with confirmed diagnosis

of metastatic prostate carcinoma (Horoszewicz et al. 1980). These cells are responsive to 5-

alpha-dihydrotestosterone (DHT) which means they contain a functioning androgen receptor

(AR) and are indicative of an early stage prostate cancer cell line even though they are

metastatic in nature (Horoszewicz et al. 1983). The cells do not grow into a uniform monolayer

but grow in clusters that must be broken apart when subculturing. They attach very lightly to

the substrate plastic, will rapidly acidify the media and are a hypertetraploid. They contain

high amounts of excess chromosome numbers with 22% of the cells containing 84

chromosomes, while cells with 86 (20%) and 87 (18%) also occur at high frequencies in culture.

Two forms of androgen receptor exist in LNCaP cells, a 110kDa and 112kDa with the 112kDa

shown to be phosphorylated form (Alimirah et al. 2006).

Page 73: Investigating the effects of, the isoflavone, Phenoxodiol

53

3.2.3.1 LNCAP MEDIA FORMULATION

LNCaP cells (cat# CRL-1740) were sourced from the American Type Culture Collection (ATCC)

and grown in a media comprising of RPMI 1640 (Gibco Cat#21870) supplemented with; 10%

fetal bovine serum (FBS, DKSH Australia Cat#FBS S101), 1% penicillin streptomycin (Gibco

Cat#15070), 1% Glutamax™ (Gibco Cat#35050), 10mM HEPES buffer (Sigma cat#H4034), 4.5g/L

d-Glucose (Sigma cat#G7528) and 1mM sodium pyruvate (Sigma cat#S8636).

3.2.4. DU145 CELLS

DU145 cells were isolated by (Stone et al. 1978) from a grade IV brain metastasis of prostate

cancer in a 69 year old Caucasian male. They are used as a “classical” example of late stage

prostate cancer and have moderate metastatic potential compared to PC3 cells. The cell line is

only weakly positive for acid phosphatase and exhibits very low DHT activity and is considered

androgen receptor (AR) negative. (Alimirah et al. 2006) have since determined that DU145 and

PC3 cells contain AR in measurable levels however, the cells do not respond to AR stimulation

and therefore have a broken pathway. Although they are not truly AR negative as they do

express a form of the receptor, it is a completely inactive form. DU145 cells are a hypotriploid

human cell line.

3.2.4.1 DU145 MEDIA FORMULATION

DU145 cells (cat#HTB-81) were sourced from the ATCC and grown in media comprising of

Minimum Essential Medium Earle’s (MEM, Gibco cat#11090) supplemented with; 10% fetal

bovine serum, 1% penicillin streptomycin, 1% Glutamax™, 1mM sodium pyruvte and 100µM

MEM non-essential amino acids solution (Gibco cat#11140).

Page 74: Investigating the effects of, the isoflavone, Phenoxodiol

54

3.2.5. PC3 CELLS

PC3 cells were isolated by (Kaighn et al. 1979) from a grade IV bone metastasis of prostate

cancer in a 62 year old Caucasian male. They are used as a “classical” example of late stage

prostate cancer and exhibit high metastatic potential. The cell line is only weakly positive for

acid phosphatase and exhibits very low DHT activity and is considered androgen receptor (AR)

negative. (Alimirah et al. 2006) have since determined that DU145 and PC3 cells contain AR in

measurable levels but that the cells do not respond to AR stimulation and therefore have a

broken pathway, but are not truly AR negative as they do express a form of the receptor, just

an inactive form.

3.2.5.1 PC3 MEDIA FORMULATION

PC3 cells (cat#CRL-1435) were sourced from ATCC and grown in media comprising of F-12K

(Kaighns Modification, Gibco cat#21127) supplemented with; 10% fetal bovine serum, 1%

penicillin streptomycin and 1% Glutamax™.

3.3. DRUG DILUTIONS

3.3.1. PHENOXODIOL

151mg of Phenoxodiol, 2H-1-benzopyran-7-0,1,3-[4-hydroxyphenyl], was supplied from

Novogen Pty Ltd in pink, slightly sticky, powdered form, in an amber glass bottle and was

stored at 4°C in a container with desiccation crystals to prevent moisture build up. Phenoxodiol

was delivered into the cells with DMSO as the carrier vehicle solution; hence all treatments

Page 75: Investigating the effects of, the isoflavone, Phenoxodiol

55

include a vehicle control group and the 10µM concentration has extra carrier added to

equilibrate vehicle control between treatment groups.

3.3.2. PHENOXODIOL WORKING STOCK

Phenoxodiol has a molecular weight of 240.26g/L and the treatment concentrations used in

this thesis were determined to be 10µM and 30µM by Novogen. To get the dose to this level

PXD was diluted at 10mg/mL, in cell culture sterile DMSO, which gave a concentration of

approximately 41600µM.

EQUATION 1 PHENOXODIOL WORKING STOCK EQUATION

3.3.3. PHENOXODIOL WORKING SOLUTION

Following dilution in DMSO, PXD working stock was stored in a covered vial at -20°C for a

maximum of three weeks before being discarded. Phenoxodiol was further diluted to make

working solutions in the complete growth media appropriate to each cell line. Combining 10µL

PXD stock (41600µM) with 990µL media resulting in a working solution with a concentration of

416µM which was then further diluted to reach the final cell concentrations of 10µM and

30µM. Vehicle control was standardised between all samples by addition of appropriately

Page 76: Investigating the effects of, the isoflavone, Phenoxodiol

56

concentrated DMSO in media to make DMSO working solution 10µL of DMSO was added into

990µL appropriate growth media per cell line. Where any dilution was performed a mirror

dilution containing only DMSO was performed to get an equally concentrated vehicle control

solution.

TABLE 3 PHENOXODIOL TREATMENT CALCULATIONS PER ML OF MEDIA

Treatment Group DMSO Working

Solution µL

Phenoxodiol

Working Solution µL

Media µL

DMSO Control 72.15µL - 927.75µL

10µM Phenoxodiol 48.11µL 24.04 µL 927.75µL

30µM Phenoxodiol - 72.15µL 927.75µL

EQUATION 2 PHENOXODIOL WORKING SOLUTION EQUATION

Page 77: Investigating the effects of, the isoflavone, Phenoxodiol

57

3.3.4. PHENOXODIOL TREATMENT

Cells were seeded at the appropriate level and grown for 48 hours. After 48 hours growth

media was aspirated from cells and replaced with appropriate treatment of Phenoxodiol;

Vehicle Control (DMSO), 10µM PXD and 30µM PXD. Following 24 or 48 hours of treatment,

media was removed and centrifuged to retain floating cells, and the experiment carried out.

3.3.5. DOCETAXEL

100mg of Docetaxel, 1,7β, 10β-trihydroxy-9-oxo-5β, 20-epoxytax-11-ene-2α, 4, 13α-triyl 4-

acetate-2-benzoate-13-{(2R,3S)}-3-[(tert-butoxycarbonyl)amino]-2-hydroxy-3-

phenylpropanoate}, was purchased (Sigma Aldrich Cat#01885) in a white crystalline powdered

form. Powder was supplied in an amber glass bottle and was stored at -20°C in a container

with desiccation crystals to prevent moisture build up. Docetaxel was delivered into the cells

with DMSO as the carrier vehicle solution in two separate working solutions. All treatments

include two vehicle control groups to equilibrate vehicle control between treatment groups.

3.3.6. DOCETAXEL WORKING STOCK AND WORKING SOLUTION

The following dilution was performed on Docetaxel to produce the two usable working

solutions (1.2378µM and 123.78nM) for use in an isobologram study with Phenoxodiol.

Docetaxel has a molecular weight of 807.88g/L and the concentrations determined for use in

this thesis were 0.1, 1, 5, 10 and 100nM. Docetaxel working solutions were discarded after use

while working stock was kept at -20°C in a covered, sealed, glass bottle for a maximum of two

weeks before being discarded. Where any dilution was performed a mirror dilution containing

only DMSO was performed to get an equally concentrated vehicle control solution.

Page 78: Investigating the effects of, the isoflavone, Phenoxodiol

58

EQUATION 3 DOCETAXEL STOCK AND WORKING SOLUTIONS

3.4. OPTIMISATION OF PROLIFERATION

MTS proliferation assays were performed to determine appropriate cell seeding rates for 96

well, 6 well and 10cm plate experiments over 48 hours of growth and 48 hours of treatment.

The CellTiter 96® AQueous One Solution Cell Proliferation Assay (Promega cat#G3580) is a

colorimetric method for determining the number of viable cells in proliferation, cytotoxicity or

chemosensitivity assays. The CellTiter 96® AQueous One Solution Reagent contains a tetrazolium

compound 3-(4,5-dimethylthiazol-2-yl)-5-(3-carboxymethoxyphenyl)-2-(4-sulfophenyl)-2H-

tetrazolium (MTS) and an electron coupling reagent phenazine ethosulfate (PES). PES has

enhanced chemical stability, which allows it to be combined with MTS to form a stable

solution. MTS absorbance was measured at 492nm on a Labsystems Multiskan RC plate

reader.

3.4.1. DETERMINATION OF CELL SEEDING CONCENTRATIONS

To determine the appropriate number of cells to use in 96 and 6 well plate assays an initial cell

seeding concentration assay was performed, on all three cells lines, to prevent cells from

Page 79: Investigating the effects of, the isoflavone, Phenoxodiol

59

reaching crowding induced senescence at the end of the treatment period (96 hours total

growth). Cells were seeded into clear 96 well plates (Sarstedt cat#83.1835, Corning

cat#CLS3595) with concentrations of; 500, 1000, 1500, 2000, 2500, 3000, 4000 and 5000

cells/100µL. Cells were seeded in the opposite direction to the treatment groups, if the

treatments are in the direction of wells 2-3-4-5 then the wells are seeded B2-C2-D2-E2 and

vice versa, this prevented bias in treatment groups caused by unequal cell volumes due to

poor mixing. After 48 hours of growth all wells had the media aspirated and replaced with

100µL of fresh complete cell media. Plates underwent MTS proliferation analysis after 6, 24

and 48 hours of total growth post initial 48 hour period. Assays were performed by adding a

20µL volume of the CellTiter 96® AQueous One Solution Reagent directly to the 100µL of sample

media and cells, incubating for 3 hours at 37°C 5% CO2, and then recording absorbance at

492nm with a Labsystems Multiskan RC plate reader. The quantity of formazan product as

measured by the amount of 492nm absorbance is directly proportional to the number of living

and metabolising cells in culture.

TABLE 4: SEEDING CONCENTRATIONS AND VOLUMES

Cell Line 96 Well Plate (100µL) 6 Well plate (2ml) 10cm Dish (5ml)

LNCaP 30,000 cells/ml 54,000 cells/ml 100,000 cells/ml

DU145 18,000 cells/ml 32,000 cells/ml 100,000 cells/ml

PC3 25,000 cells/ml 45,000 cells/ml 100,000 cells/ml

3.4.2. PHENOXODIOL MTS ASSAY

Phenoxodiol (PXD) was shown to increase activity of the MTS assay product formazen without

cellular interface, all PXD MTS proliferation assays included n=2 no cell controls for all

treatment groups to remove background absorbance increase caused by PXD. In final analysis

Page 80: Investigating the effects of, the isoflavone, Phenoxodiol

60

no cell controls were averaged and removed from each of the cell samples which varied from

n=4 to n=6. Experimentation began with cells seeded at appropriate rates into clear 96 well

plates (Sarstedt#83.1839, Corning#CLS3516) with a total volume of 100µL per well. All assays

included the following groups; Vehicle control (DMSO), 10µM PXD and 30µM PXD over 24 and

48 hours. Cells were grown for 48 hours in an incubator at 37°C 5% CO2 before the complete

cell media was aspirated and replaced with 100µL of appropriate treatment media then left to

grow at incubator at 37°C 5% CO2 for 24 and 48 hours. Following appropriate growing period

20µL of CellTiter 96® AQueous One Solution was added to each well, mixed gently with a pipette

then left for 3 hours in an incubator at 37°C 5% CO2 then analysed on a Labsystems Multiskan

RC plate reader at 492nm.

Vehicle

Control

10µM

PXD

30µM

PXD

FIGURE 10 MTS ASSAY SETUP

MTS assay setup with cell treatments indicated by yellow and no cell controls indicated by red,

multiple cell lines could be grown on the same plate for 24 or 48 hour treatment however the

two time points could not be performed on the same plate due to risk of contamination.

Page 81: Investigating the effects of, the isoflavone, Phenoxodiol

61

3.4.3. PHENOXODIOL DOCETAXEL ISOBOLOGRAM

An isobologram is a test for properties of interference, additivity or synergy between two or

more chemicals, and was performed between phenoxodiol and the common prostate cancer

treatment Docetaxel, a taxane based anti-microtubulin chemotherapeutic agent. The

isobologram had 24 individual treatment groups with 4 Phenoxodiol (0,5,10,30µM PXD) and 6

Docetaxel (and 0,0.1,1,5,10,100nM DOC) treatment concentrations, compared to each other.

All treatments had a cell n=4 and no cell control of each n=2 resulting in a total of 144 wells

per cell line per time course. Due to this high number only the 48 hour isobologram was

performed to allow for maximal treatment variation. As the outside rows of 96 well plates

have been shown to have a statistically significant difference in light absorption they were

excluded from use resulting in the need for a single plate for each Phenoxodiol treatment

group (4 plates total used). Cells were seeded at appropriate rates in 100µL complete growth

media and left for 48 hours.

Page 82: Investigating the effects of, the isoflavone, Phenoxodiol

62

DOC

Vehicle

Control

0nM

DOC

0.1nM

DOC

5nM

DOC

10nM

DOC

100nM

DOC

FIGURE 11 96 WELL ISOBOLOGRAM SETUP

The 96 well plate setup required for each cell line and Phenoxodiol concentration with yellow

indicating cell containing wells and red indicating no cell control wells. Each cell line had 4

plates with varying levels of Phenoxodiol used; Vehicle Control (DMSO), 5µM, 10µM and 30µM

Phenoxodiol. Cells were treated over a 48 hour period with all treatment groups balanced to

the most concentrated DMSO group (30µM PXD + 100nM DOC).

Following 48 hours of growth cells had their media aspirated and an appropriate media added

to each well. Media was made to control for the DMSO vehicle used to deliver both the PXD

and DOC treatments. In media preparation two sets of DOC working solutions were required

(1.2378µM and 123.78nM) to be able to accurately pipette the volumes required across the

spread of concentrations and as such two vehicle (DMSO) solutions were also required.

Page 83: Investigating the effects of, the isoflavone, Phenoxodiol

63

TABLE 5 ISOBOLOGRAM SETUP FOR VEHICLE CONTROL AND TREATMENT GROUPS

PXD and DOC

Concentrations

(µM and nM)

PXD

Vehicle

Control

(DMSO)

(µL)

PXD

Working

Solution

(µL)

(416µM)

DOC

Vehicle

Control

1

(DMSO)

(µL)

DOC

Working

Solution

1 (µL)

(1.2378

µM)

DOC

Vehicle

Control

2

(DMSO)

(µL)

DOC

Working

Solution

2 (µL)

(123.78

nM)

Media

to total

1mL (µL)

P 0 D 0 72.11 - 80.79 - 8.08 - 839.02

P 0 D 0.1 72.11 - 80.79 - 7.27 .81 839.02

P 0 D 1 72.11 - 80.79 - - 8.08 839.02

P 0 D 5 72.11 - 76.75 4.04 8.08 - 839.02

P 0 D 10 72.11 - 72.71 8.08 8.08 - 839.02

P 0 D 100 72.11 - - 80.79 8.08 - 839.02

P 5 D 0 60.09 12.02 80.79 - 8.08 - 839.02

P 5 D 0.1 60.09 12.02 80.79 - 7.27 .81 839.02

P 5 D 1 60.09 12.02 80.79 - - 8.08 839.02

P 5 D 5 60.09 12.02 76.75 4.04 8.08 - 839.02

P 5 D 10 60.09 12.02 72.71 8.08 8.08 - 839.02

P 5 D 100 60.09 12.02 - 80.79 8.08 - 839.02

P 10 D 0 48.07 24.04 80.79 - 8.08 - 839.02

P 10 D 0.1 48.07 24.04 80.79 - 7.27 .81 839.02

P 10 D 1 48.07 24.04 80.79 - - 8.08 839.02

P 10 D 5 48.07 24.04 76.75 4.04 8.08 - 839.02

P 10 D 10 48.07 24.04 72.71 8.08 8.08 - 839.02

P 10 D 100 48.07 24.04 - 80.79 8.08 - 839.02

P 30 D 0 - 72.11 80.79 - 8.08 - 839.02

P 30 D 0.1 - 72.11 80.79 - 7.27 .81 839.02

P 30 D 1 - 72.11 80.79 - - 8.08 839.02

P 30 D 5 - 72.11 76.75 4.04 8.08 - 839.02

P 30 D 10 - 72.11 72.71 8.08 8.08 - 839.02

P 30 D 100 - 72.11 - 80.79 8.08 - 839.02

The total number of solutions required were two DOC working and vehicle control solutions,

one PXD working solution and one PXD vehicle control solution. Following 48 hours of

Page 84: Investigating the effects of, the isoflavone, Phenoxodiol

64

treatment, with PXD and DOC, 20µL of CellTiter 96® AQueous One Solution Reagent was added to

each well and mixed gently then incubated at 37°C 5% CO2 for 3 hours before being analysed

on a Labsystems Multiskan RC plate reader at 492nm.

3.4.4. PHENOXODIOL AND CASPASE INHIBITION

In order to understand potential interactions between the caspase signalling pathway and

Phenoxodiol an experiment was performed using the pan caspase inhibitor Z-VAD-FMK

(Carbobenzoxy-valyl-alanyl-aspartyl-(O-methyl)-fluoromethylketone) purchased from BD

Pharmingen Cat#550377. Z-VAD-FMK is a cell permeable general caspase inhibitor that

irreversibly binds to the catalytic site of caspase proteases and inhibits apoptosis. 1mg of Z-

VAD-FMK was reconstituted in 214µL of DMSO with a resulting stock concentration of 10mM.

A final concentration of 10µM Z-VAD-FMK was used in the study with the maximal PXD dose of

30µM over a 48 hour period. Before use a titration was performed indicating that 100µM Z-

VAD-FMK was cytotoxic and 100nM Z-VAD-FMK was not sufficiently concentrated to prevent

caspase activation from a UV irradiation source (measured by analysing activated Caspase-3

post irradiation). A final concentration of 10µM Z-VAD-FMK was chosen as it prevented

Caspase-3 from activating post UV irradiation however cells did eventually enter into cell death

when measured after 48 hours. A 100 fold dilution in complete cell media was performed to

get a working solution of 100µM Z-VAD-FMK. Cells were seeded at appropriate levels in clear

96 well plates with cells n=4 and no cell controls n=2 for the treatment groups; Vehicle control,

10µM Caspase Inhibitor, 30µM PXD and 10µM CI + 30µM PXD. Following 48 hours of growth in

a 37°C 5% CO2 incubator the media was aspirated and replaced with the appropriate

treatment media.

Page 85: Investigating the effects of, the isoflavone, Phenoxodiol

65

TABLE 6 CASPASE INHIBITION AND PHENOXODIOL TREATMENT SOLUTIONS

Treatment

Group

Phenoxodiol

DMSO

Working

Solution (µL)

Phenoxodiol

Working

Solution

416µM (µL)

Caspase

Inhbitor

DMSO

Working

Solution (µL)

Caspase

Inhbitor

Working

Solution

100µM (µL)

Media µL to

make 1mL

DMSO

Control

72.15 - 100 - 827.85

10µM

Caspase

Inhibitor (CI)

72.15 - - 100 827.85

30µM

Phenoxodiol

- 72.15 100 - 827.85

10µM CI +

30µM

Phenoxodiol

- 72.15 - 100 827.85

Following 48 hours of treatment, with PXD and CI, 20µL of CellTiter 96® AQueous One Solution

Reagent was added to each well and mixed gently then incubated at 37°C 5% CO2 for 3 hours

before being analysed on a Labsystems Multiskan RC plate reader at 492nm.

3.4.5. PHENOXODIOL AND DOCETAXEL COMBINATION TREATMENT

In order to further understand the interaction of phenoxodiol and docetaxel an experiment

was performed to analyse pre-treatment of cells with phenoxodiol and subsequent cell

cytotoxicity when both solutions were recombined. Cells would be exposed to either; 48 hours

of 10µM or 30µM phenoxodiol only, 48 hours of 100nM docetaxel only, 48 hours of

10µM/30µM phenoxodiol 100nM docetaxel combination, 48 hours of 10µM/30µM

phenoxodiol and 24 hours of docetaxel or 48 hours docetaxel and 24 hours of 10µM/30µM

Page 86: Investigating the effects of, the isoflavone, Phenoxodiol

66

phenoxodiol. Cells were seeded at appropriate rates in 100µL complete growth media and left

for 48 hours with an n=4 for treatment groups and n=2 for no cell (media only) controls. Media

was then aspirated off and replaced with treatment media (as seen in table 4) with the two 48

hour/24 hour combinations receiving a normal dose of either; 10µM/30µM phenoxodiol or

100nM docetaxel. All treatments were controlled for DMSO vehicle concentration. After 24

hours of treatment the final treatment was given in concentrated form to the treatment wells

that required it, either 100nM docetaxel to the 48 hour PXD 24 hour DOC treatment or

10µM/30µM PXD to the 48 hour DOC 24 hour PXD treatment, again DMSO vehicle was

controlled for at this point. After a further 24 hour incubation, 20µL MTS was added to each

well and incubated in the dark at 37°C 5% CO2 for 3 hours before being analysed on a

Labsystems Multiskan RC plate reader at 492nm.

3.4.6. PHENOXODIOL AND PURIFIED SFRP4 PROTEIN COMBINATION

TREATMENT

In order to assess the potential interaction of Phenoxodiol with the canonical Wnt/β-catenin

pathway and non-canonical Wnt pathway, an experiment was performed utilising the frizzled

receptor antagonist secreted frizzled related protein 4 (sFRP4). Cells would be proliferated for

48 hours then exposed to 48 hours of treatment utilising; vehicle control (PBS+DMSO),

125pg/mL sFRP4, 250ph/mL sFRP4, 500pg/mL sFRP4, 10µM Phenoxodiol, 30µM Phenoxodiol

and a combination of 500pg/mL sFRP4/30µM Phenoxodiol. Cells were seeded at appropriate

rates in 100µL complete growth media and left for 48 hours with an n=4 for treatment groups

and n=2 for no cell (media only) controls. Following 48 hours of growth the complete media

was aspirated off and replaced with treatment media (below) with the working solution

concentration of sFRP4 purified protein being 125µg/mL in PBS. After 48 hours of incubation,

Page 87: Investigating the effects of, the isoflavone, Phenoxodiol

67

20µL MTS was added to each well and incubated in the dark at 37°C 5% CO2 for 3 hours before

being analysed on a Labsystems Multiskan RC plate reader at 492nm.

TABLE 7 SFRP4 & PHENOXODIOL TREATMENT SOLUTIONS

Treatment

Group

DMSO

Working

Solution (µL)

Phenoxodiol

Working

Solution

416µM (µL)

PBS (µL) sFRP4

Working

Solution

(125µg/mL)

Media µL to

1mL

Vehicle

Control

(DMSO + PBS)

72.15 - 4 - 923.85

125pg/mL

sFRP4

72.15 - 3 1 923.85

250pg/mL

sFRP4

72.15 - 2 2 923.85

500pg/mL

sFRP4

72.15 - - 4 923.85

10µM

Phenoxodiol

48.11 24.04 4 - 923.85

30µM

Phenoxodiol

- 72.15 4 - 923.85

500pg/mL

sFRP4 + 30µM

Phenoxodiol

- 72.15 - 4 923.85

3.5. REACTIVE OXYGEN SPECIES DETECTION

Reactive oxygen species (ROS), specifically nitric oxide (NO), detection was performed using a

Griess Reagent System (Promega cat#G2930). The Griess system uses sulfanilamide and N-1-

naphthylethylenediamine dichloride (NED) under acidic conditions supplied by phosphoric

acid. This assay measures the NO2- which is one of two stable non-volatile breakdown products

from NO production.

Page 88: Investigating the effects of, the isoflavone, Phenoxodiol

68

Cells were seeded at the appropriate rate on a 96 well plate with cell controls of n=4 and no

cell controls n=2 with 4 treatment groups; Vehicle control, 10µM PXD, 30µM PXD, DEAN and

SNP. Following 48 hours of growth cells were treated with appropriate levels of Vehicle

control, Phenoxodiol or DEAN for 24 and 48 hours. DEAN (Diethylamine NONOate

diethylammonium salt, Sigma cat#D5431) a nitric oxide stimulator was used as a positive

control to show that the cells can produce NO with appropriate stimulation. DEAN was in a

stock concentration of 100mM and was diluted 1:1000 in complete cell media to become a

final concentration of 100µM. Sodium Nitroprusside (SNP) was another positive control

utilised to indicate the cells were capable of producing nitric oxide when stimulated and was in

a 10mM stock concentration that was diluted 1:100 in complete media to a final concentration

of 100µM.

Before the assay was run, 50µL of media was removed from each well leaving a volume of

50µL. A nitrite standard reference curve was performed on each assay plate. 1mL of 100µM

nitrite solution was produced by diluting the 0.1M nitrite solution supplied with the assay kit

1:10,000 in complete media. 50µL media was added to rows B10-H10 and B11-H11 and 100µL

of prepared 100µM nitrite solution was added to A10 and A11. Using a 200µL pipette set to

50µL each row had 50µL of sample taken and added into the next row, serially diluting the

solution down until the final row (H) where the product was discarded. The standard curve

wells were left with 50µL of nitrite and media solution in all wells with a concentration scale of

100, 50, 25, 12.5, 6.25, 3.13, 1.56 and 0µM nitrite in rows A-H respectively.

The sulfanilamide and NED solutions were allowed to come up to room temperature for 15

minutes, after being stored at 4°C. 50µL of sulfanilamide solution was added to each sample

and standards well, bringing the total volume per well to 100µL and the plate was covered in

Page 89: Investigating the effects of, the isoflavone, Phenoxodiol

69

foil to protect it from the light, and incubated at room temperature for 10 minutes. Following

incubation a further 50µL of NED solution was added to each well and the plate was again

incubated and covered at room temperature for 10 minutes. Following this final incubation the

plate was analysed on a Labsystems Multiskan RC plate reader at 540nm with increases in

absorbance expression indicating increased NO output.

3.6. ACIDITY ANALYSIS

Media was analysed post treatment with Phenoxodiol to test whether the drug was altering

the acidity of the media. A pH meter was calibrated with buffer solutions before use and at a

room temperature of 24°C. Post 48 hour treatment media was placed under the probe to

analyse any pH changes between vehicle control and treatment groups.

3.7. APOPTOSIS ANALYSIS ASSAYS

Following treatment with Phenoxodiol apoptosis and necrosis levels were determined through

five assays; 3’-End Labelling DNA Fragmentation Analysis, Annexin-V-FLUOS/PI Flow

Cytometry, Sybr Gold DNA Fragmentation Analysis, JC-1 Mitochondrial Potential Detection and

Activated Caspase-3 Activity. DNA extraction was required for both 3’-End labelling and Sybr

Gold DNA fragmentation assays.

3.7.1. DNA EXTRACTION

Cells were seeded in 6-well plates, n=4, at appropriate seeding rates, and grown for 48 hours.

Following 48 hours of growth the media was removed and replaced with 2mL of treatment

Page 90: Investigating the effects of, the isoflavone, Phenoxodiol

70

media containing; Vehicle control (DMSO), 10µM PXD or 30µM PXD. After 24 and 48 hours

treatment, the media was removed, placed in a 2mL eppendorf tube and centrifuged at 800g.

The media was then discarded leaving cells behind, while 1mL of 1xPBS was placed on each cell

well. Wells were scraped with a cell scraper (Sarstedt cat#83.1832) and pipetted into the 2mL

eppendorf tube that had their media centrifuged in. Cells were again centrifuged at 800g and

1xPBS supernatant discarded. Cells were then homogenised in 1mL of DNA homogenisation

buffer4. The homogenate then had 62.5µL of 10% SDS added, was vortexed and incubated at

65°C for 30 minutes. 175µL of 8M potassium acetate was then added, vortexed and the

homogenate incubated on ice for a further 60 minutes. Samples were then centrifuged for 10

minutes at 4°C at 12,000rpm in an eppendorf refrigerated centrifuge (Eppendorf cat#5417R)

and the supernatant was transferred to a new tube. 2µL of RNaseA 500µg/mL (Roche Applied

Science cat#10109169001) was added and the mixture then incubated at 37°C for 1 hour.

Equal volumes (500µL) of phenol:chloroform/isoamyl alcohol (1:1) were added and vortexed

vigorously, then centrifuged at for 5 minutes at 4°C at 6000rpm.

Upper aqueous phase was collected and 0.1 (50µL) volumes of 3M Sodium Acetate and 2.5

volumes (1250µL) of cold 100% ethanol were added and precipitated overnight at -80°C.

Following precipitation DNA was centrifuged for 30 minutes at 14,000rpm and the supernatant

removed. The pellet was washed with 100µL of 70% ethanol by vortexing and centrifuging for

3 minutes at 14,000rpm. Ethanol was discarded and the pellet air dried for at least 1 hour in a

class I hood then resuspended in 20µL of dH2O, vortexed and placed in a 37°C incubator for 15

minutes. Solution was vortexed and gently centrifuged to ensure dissolution of DNA pellet into

the dH2O, then a 2µL sample was taken and diluted in 48µL of dH2O for spectrophotometer

analysis on a Nanodrop ND-1000 spectrophotometer (Nanodrop cat#ND-1000) with samples

outside of a 1.8-2 260/280nm range discarded. Samples were stored in a freezer at -80°C.

4 DNA Homogenisation Buffer: 0.1M NaCl, 0.01M EDTA pH 8.0, 0.3M Tris-HCl, pH 8.0, 0.2M Sucrose in

H2O

Page 91: Investigating the effects of, the isoflavone, Phenoxodiol

71

3.7.2. 3’-END LABELLING DNA FRAGMENTATION ANALYSIS

Three prime end labelling (3’-end labelling) was used as a method of quantitative and

qualitative DNA fragmentation analysis. The radioactive phosphate isotope (P32) was bound to

ddATP ([α32P]-ddATP, Amersham cat#PB10233-250uCI) and then an enzyme, terminal

transferase, tagged all fragmented DNA with the isotope/ddATP mix, which was then exposed

to film allowing for a qualitative measure of apoptosis versus necrosis. This was followed by

radiation detection on a Packard Tri-Carb 1500 Liquid Scintillation Counter, allowing a

quantitative measurement of apoptosis.

To perform a 3’-end labelling fragmentation assay DNA was isolated and quantified as

previously described. A total of 1µg of DNA was used per sample (n=4 per treatment) and the

DNA solution diluted to 29µL total volume with ddH2O then added to a 3’-End Labelling

reaction solution.

TABLE 8 3'-END LABELLING REACTION MIXTURE

Reagent Volume

DNA (1µg in 29µL H2O) 29µL

5X Reaction Buffer5 10µL

CoCl2 (25mM) solution 5µL

[α32P]-ddATP (50µCi; 3.4pmol/µL) 5µL

Terminal Transferase (25 units/µL in 50% glycerol stock diluted 80 fold in water) 1µL

Once the solution was complete samples were; vortexed briefly, quickly centrifuged and then

incubated at 37°C for 60 minutes. The reaction was terminated with the addition of 5µL of

0.25M EDTA and then 2µL of transfer RNA solution (25mg/mL) was added with 12µL 10M

ammonium acetate and 180µL of -20°C 100% ethanol before the solution was vortexed and

5 Terminal Transferase Reaction Buffer (5X): 1M Potassium Cacodylate, 0.125M Tris-HCl, 1.25mg/mL

BSA; pH 6.6

Page 92: Investigating the effects of, the isoflavone, Phenoxodiol

72

precipitated at -80°C for 60 minutes. After the DNA was precipitated samples were centrifuged

at 10,500rcf (Eppendorf Cat#5415C) for 20 minutes and the hot supernatant was discarded

before the pellet was resuspended in 55µL of 1x TE. The precipitation step containing

ammonium acetate and ethanol was repeated with the pellet air dried for 20 minutes, after

the supernatant was discarded, then diluted in 40µL of 1x TE and stored at -20°C overnight.

A 2% agarose-1x TAE gel was prepared and set while samples were thawed and 20µL of sample

combined with 4µL of DNA loading buffer6 and loaded into the gel with the excess solution

stored at -20°C. Samples underwent electrophoresis at 60 volts for between 2.5 and 4 hours in

1x TAE buffer. Following electrophoresis gels were placed into a gel slab dryer for 2 hours or

until the gel had over 90% of it’s liquid content removed. The thin gel was then placed into

plastic wrap and exposed to X-ray film (Kodak) at -80°C for 6-24 hours depending on sample

run.

Following film exposure the gel lanes were cut into individual strips and loaded into

scintillation count tubes with 2mL of liquid scintillant. Tubes were then analysed for

radioactivity using a Packard Liquid Scintillation counter and these counts per minute used to

determine the level of apoptotic fragmentation in samples.

3.7.3. ANNEXIN-V-FLUOS PROPIDIUM IODIDE FLOW CYOMETRY

An early feature of apoptosis is the translocation of the phosphatidylserine (PS) receptor from

the inner to the outer surface of the plasma membrane. Annexin-V-Fluorescein (Annexin-V-

FLUOS, AVF) is a fluorescence binding protein with a high affinity for externalised

6 DNA Loading Buffer: 30% glycerol, 0.25% bromophenol blue, 0.25% xylene cylanol

Page 93: Investigating the effects of, the isoflavone, Phenoxodiol

73

phosphatidylserine receptor, and therefore a good detector of apoptotic cells. However, AVF

can also enter necrotic/late stage apoptotic cells and bind to the PS exposed on the inner cell

membrane. By staining simultaneously with Propidium Iodide (PI), a DNA stain that fluoresces

once intercalated into double stranded DNA but is impermeable to complete cell membranes,

we can distinguish apoptotic from necrotic cells as apoptotic cells will actively exclude PI until

such a time as their membrane becomes breached. Annexin-V-FLUOS, Propidium Iodide

double staining is able to detect healthy, apoptotic and late stage apoptotic/necrotic cells and

was performed using the Roche Applied Science Annexin-V-FLUOS detection kit cat#1 828 681.

TABLE 9 CELL STAINING WITH ANNEXIN-V/PROPIDIUM IODIDE (AV/PI) DOUBLE STAIN

Cell Type AVF Staining PI Staining

Normal Cells Negative Negative

Necrotic Cells Positive Positive

Apoptotic Cells Positive Negative

Cells were seeded in 6-well plates, n=4, at appropriate seeding rates, and grown for 48 hours

with an extra set of 3 wells required for machine setup. Following 48 hours of growth the

media was removed and replaced with 2mL of treatment media containing; Normal Media

(Setup Cells), Vehicle control (DMSO), 10µM PXD or 30µM PXD. After 24 and 48 hours

treatment, the media was removed, placed in a 2mL eppendorf tube and centrifuged at 400g

for 5 minutes. Cells had 1mL of 1xPBS added, then aspirated and centrifuged at 400g for 5

minutes, before 500µL 0.5% Trypsin-EDTA solution was added to each well and the cells

incubated at 37°C 5% CO2 for 5 minutes. 200µL of complete media containing FBS was then

added to inhibit further trypsin activity. Cells were then washed down the face of the 6 well

plate, washing off any stuck cells, before being centrifuged at 400g for 5 minutes. Media was

then removed from the cell pellet and pellet resuspended in 100µL of appropriate labelling

Page 94: Investigating the effects of, the isoflavone, Phenoxodiol

74

solution. Each time AV/PI staining was performed a set of setup cells was included in the

extraction and treatment run, grown in normal media. These cells allowed for analysis

between runs and correct FACS machine setup.

TABLE 10 CELL STAINING SETUP FOR AV/PI

Cell Type Stain

Control Cells (n=1) No AVF and No PI

Control Cells (n=1) AVF Only

Control Cells (n=1) PI Only

Vehicle Control Cells (n=4) AVF and PI

Treatment Cells (n=4) AVF and PI

TABLE 11 AV/PI LABELLING SOLUTIONS

Staining Buffers Labelling Solution Recipe

No AVF + No PI Added 100µl Incubation buffer to cell pellet

AVF Only Mixed 204µl Incubation buffer + 4µl Annexin-V (enough for 1 sample)

Added 100µl to cell pellet

PI Only Mixed 204µl Incubation buffer7 + 4µl Propidium Iodide8 (enough for 1

sample) Added 100µl to cell pellet

AVF + PI Mixed 1ml Incubation buffer + 100µl Annexin-V + 100µl Propidium

Iodide (enough for 10 samples) Added 100µl to cell pellet

After 100µL of appropriate buffer was added to each sample, cells were incubated for 15

minutes at 25°C in darkness. Incubation buffer was added with volume depending on cell

pellet size, between 200 and 400µL was used to ensure an suitable event rate through the

7 Incubation Buffer: 10mM HEPES/NaOH, pH 7.4, 140mM NaCl, 5mM CaCl2 (keeps for 3 months at 4°C)

8 PI Solution: Stock solution 50µg/ml in dH2O (PI is light sensitive and was stored at 4°C in the dark)

Page 95: Investigating the effects of, the isoflavone, Phenoxodiol

75

FACSCalibur (BD Biosciences cat#342976) at low speed. The flow cytometer was configured

using a 488nm excitation and a 515nm bandpass filter for fluorescein detection and a 600nm

filter for PI detection. Electronic compensation of the instrument was required to exclude

overlapping of the two emission spectra and was performed by the FACS specialist, Dr Kathy

Heel from the Centre for Microscopy, Characterisation & Analysis, University of Western

Australia. Post experimental graphing was performed using FlowJo software version 7.2.5,

engine 1.999995. Once cell populations were gated appropriately to exclude

fragments/doublets/triplets the cells fell within the 3 areas, No staining (live cells), Annexin-V-

FLUOS only (Early apoptotic cells) and double staining of Annexin-V-FLUOS and Propidium

Iodide (Late stage apoptosis/Necrotic death). The % of cells within each set of gates was then

graphed.

3.7.4. SYBR GOLD FRAGMENTATION ANALYSIS

SYBR Gold staining was performed as a measure of qualitating DNA quality and checking 3’-end

fragmentation laddering. SYBR Gold (Invitrogen cat#S11494) is a DNA intercalating agent that

allows for detection of DNA in a cell. Apoptosis produces a characteristic laddering effect

within the cellular DNA due to initiation of endogenous endonuclease activity on sites between

nucleosomes. Nucleosomes lie approximately 200 base pairs (bp) apart and the result of

endonuclease activation is 180bp internucleosomal fragments and multiples thereof. Analysis

of cellular DNA through an agarose gel provides a characteristic laddering effect if cells are

undergoing apoptosis, or a smear if the cells are necrotic. DNA was extracted as previously

described9 and 1µg of DNA was determined by spectrophotometer then placed into a 2%

agarose Tris-acetate EDTA (TAE) gel with 1µL of Loading Buffer10 (Promega Cat# G1881) per

9 3.7.1 DNA Extraction Page: 69

10 Loading Buffer: 0.4% orange G, 0.03% bromophenol blue, 0.03% xylene cyanol FF, 15% Ficoll® 400,

10mM Tris-HCl (pH 7.5) and 50mM EDTA (pH 8.0)

Page 96: Investigating the effects of, the isoflavone, Phenoxodiol

76

5µL of sample. 40 volts was applied to the gel over a period of 4 hours to allow for high

resolution of the samples before the gel was placed into a container with 1xSYBR Gold11 TAE12

buffer covering the gel completely then placed in the dark for 40 minutes. Following staining

the gel was placed on a Kodak Imager 2000 and excited at 465nm while emission was read at

535nm.

FIGURE 12 LNCAP SYBR GOLD VISUALISED DNA FRAGMENTATION WITH DNA LADDER, LOW WEIGHT

DNA FRAGMENTATION IS VISIBLE

3.7.5. JC-1 MITOCHONDRIAL POTENTIAL ASSAY

At the onset of apoptosis, the mitochondrial membrane is rapidly depolarised (Hong et al.

2004). Accompanying this, 5,5’,6,6’-tetrachloro-1,1’,3,3’-tetraethylbenzimidazolcarbo-cyanine

11

SYBR Gold Stock 10,000x made up to 1x in TAE 12

40 mM Tris-acetate, 1 mM EDTA pH 7.5–8.0

Page 97: Investigating the effects of, the isoflavone, Phenoxodiol

77

iodide (JC-1) incorporates into mitochondria and forms monomers that fluoresce green

(520nm), whereas at high membrane potentials it forms J-aggregates which fluoresce red

(590nm) (Salvioli et al. 1997). Therefore the ratio of aggregated versus monomeric JC-1 gives a

quantitative representation of mitochondrial membrane permeability such that a low

red:green ratio is indicative of apoptosis (Zamzami et al. 2000; Zamzami et al. 2000). The JC-1

mitochondrial assay (Invitrogen, cat#M34152) assesses mitochondrial membrane potential in

live cells and allows for quantification of depolarisation of mitochondria, an event

characteristic of early stage cell death signalling. JC-1 penetrates the cytosol of eukaryotic cells

and exhibits potential-dependent accumulation in mitochondria, indicated by a fluorescence

emission shift from green (~529nm) to red (~590nm). Consequently, mitochondrial

depolarisation is indicated by a decrease in the red/green fluorescence intensity ratio. As a

positive control the kit used carbonylcyanide-4-trifluoromethoxyphenylhydrazone (FCCP) a

potent mitochondrial depolarising agent.

Cells were seeded with appropriate number onto white 96 well fluorescent plates (Greiner Bio-

One cat#675083) and following 48 hours of growth media was removed and appropriate

treatment used with cell control n=4 and no cell controls n=2. Four treatments were used with;

normal media for FCCP staining, vehicle control (DMSO), 10µM PXD and 30µM PXD over 24

and 48 treatment times. JC-1 staining solution was made up each assay by combining 2.5mM

JC-1 stock solution (Molecular Probes cat#T-3168) 1:75 in media without FBS. FCCP staining

solution was made up using 1:75 JC-1 buffer with 1:66 FCCP (50mM stock) in media without

FBS.

Following treatment with Phenoxodiol the media was removed and either 50µL JC-1 or 50µL of

JC-1/FCCP staining solution was added to each well and incubated at 37°C 5% CO2 covered to

protect the plate from light, for one hour. Following incubation staining solution was removed

and 200μl PBS with 5% Bovine Serum Albumin (BSA) was added to each well to quench excess

Page 98: Investigating the effects of, the isoflavone, Phenoxodiol

78

fluorescence. The plate was incubated for 5 minutes at 37°C 5% CO2 before the PBS/BSA

solution was removed and 100µL PBS added to each well. The assay was read on a BMG

Labtech Fluostar Optima fluorescent plate reader with the plate chamber set to 37°C and each

well measured in a 5x5 cross-section. A set gain level was used between all cell lines and time

points. Excitation/Emission was measured from the top, green fluorescence was measured at

485nm excitation and 520nm emission while red fluorescence was measured at 544nm

excitation and 590nm emission. Red results were divided by green to get the ratio of red to

green emission, with a low ratio indicated high depolarisation rate and early apoptotic

induction.

3.7.6. CASPASE-3 ACTIVITY ASSAY

Activated Caspase-3 activity was measured using the EnzChek Capase-3 Assay Kit #2 from

Invitrogen (cat#E13184) which contains a rhodamine 110 substrate compound Z-DEVD-R110

which is turned, by activated Caspase-3, into the green fluorescent rhodamine 110. Cells were

plated onto a black fluorescent plate (Greiner Bio-One cat#675086) at the appropriate level

and grown for 48 hours in normal media. Media was aspirated and replaced with treatment

media; DMSO Control, 10µM and 30µM PXD n=4 plus n=2 no cell controls.

Caspase-3 staining dye was prepared with 100µL per sample containing; 20µL 5x Reaction

Buffer13, 0.5µL Dithiothreitol (DTT), 5µL 20x Cell Lysis Buffer14, 0.5µL Z-DEVD-R110 substrate

and 74µL of deionised water. Following 24 and 48 hours of total exposure to treatment media,

media was aspirated from the cells and replaced with 100µL of Caspase-3 staining dye per

sample. 96 well plates were covered with foil and placed on ice to incubate for 30 minutes

13

Reaction buffer: 20mL of 50mM PIPES, pH 7.4, 10mM EDTA, 0.5% CHAPS 14

Cell lysis buffer: 1.5mL of 200mM TRIS, pH 7.5, 2M NaCl, 20mM EDTA, 0.2% Triton-X-100

Page 99: Investigating the effects of, the isoflavone, Phenoxodiol

79

before being incubated at room temperature for a following 30 minutes. The assay was read

on a BMG Labtech Fluostar Optima fluorescent plate reader with the plate chamber set to 37°C

and each well measured in a 5x5 cross-section.

Excitation/Emission was measured from the top with fluorescence occurring when activated

Caspase-3 had converted Z-DEVD-R110 into R110. Cells were measured at 496nm excitation

and 520nm emission with a set gain level between cell lines and time points used.

3.8. CELL CYCLE ANALYSIS

To determine if Phenoxodiol induced any potent cell cycle effects propidium iodide cell cycle

analysis was performed using a BD FACS Canto II flow cytometer (Becton Dickinson, Cat#

338962). Cells were analysed for DNA content through fluorescence of propidium iodide which

increases in fluorescence by 20 to 30 fold once intercalated into nucleic acids. Cells are

measured with a given fluorescence level and through shape and size analysis G1 content can

be attributed to a given population, G2 is assumed to be between 1.9 and 2.1 fold greater in

DNA fluorescence with S phase cells between the two population peaks. Analysis of forward

scatter (FSC) and side scatter (SSC) area (-A), height (-H) and width (-W) in the cell population

allowed for determination of cell size and cell complexity/granularity respectively. Using

complexity and cell size allowed cell fragments, doublets and triplets to be discarded and

gated out of analysis with a minimum of 10,000 individual cells within gated regions being

counted for significance to be reached.

Page 100: Investigating the effects of, the isoflavone, Phenoxodiol

80

3.8.1. CELL CYCLE PREPARATION

Cells were plated at 1x105 cell/mL with 5mL total media in 100mm plates (BD Falcon

cat#353003) and grown for 48 hours before appropriate treatment was used; Normal media,

vehicle control, 10µM and 30µM PXD (n=4). Cells were grown for 24 and 48 hours then media

was aspirated and centrifuged for 4 minutes at 400g while dishes were coated with 4mL of 1x

PBS and washed. Media was removed from the cell pellet and replaced with 1x PBS aspirated

from cell plates and centrifuged again for 4 minutes at 400g while 3mL of trypsin-EDTA was

placed into each plate and incubated for at least 5 minutes at 37°C 5% CO2.

Once trypsinisation had occurred, determined by a majority of cells separating, 3mL of 1x PBS

was added to each plate and cells washed then placed into tubes containing the cell pellet

from media and 1x PBS wash centrifugation. Cells were centrifuged for 5 minutes at 400g

,leaving a large cell pellet, before media was aspirated leaving just the pellet behind. Cells

were then gently vortexed while a 4mL, -20°C, 70% ethanol 30% water mixture was added

drop-wise into the vortexing cells causing fixation. Cells were stored at 4°C for up to 3 months

in the 70% ethanol solution with longer storage possible although not performed in this

experiment.

Following fixation, on the day of analysis, cells were centrifuged out of the 70% ethanol

solution at 800g for 5 minutes, a higher g rating is required as the ethanol partly crenates the

cells and prevents successful centrifugation at the lower 400g rating previously used. Following

centrifugation ethanol was removed and replaced with 3mL of 1x PBS before being centrifuged

at 400g for 5 minutes. PBS was aspirated and again replaced with 1x PBS and centrifuged for 5

minutes at 400g to ensure complete cell rehydration. Following second and final rehydration

step the cell had 1x PBS removed and replaced by between 400 and 1mL of propidium iodide

Page 101: Investigating the effects of, the isoflavone, Phenoxodiol

81

staining solution before being incubated for 30 minutes at 37°C to allow RNAse activity to

occur.

TABLE 12 PROPIDIUM IODIDE STAINING SOLUTION PER 30ML

Chemical Volume

Fetal Bovine Serum 600µL

Sodium Citrate 30mg

EDTA 2.3mg

Triton X-100 9µL

Propidium Iodide 1.5mg

RNAse A (10u/µL) 15µL

H2O 29376µL

3.8.2. CELL CYCLE FLOW CYTOMETRY

When cell cycle flow cytometry was performed, negative controls were used to setup the

machine to within parameters to allow for accurate gating, assessment and analysis of

samples. Samples were optimally analysed at between 100 and 200 events per second, greater

events per second were prevented by addition of extra PI buffer and low levels were

prevented by using low initial volumes of buffer to stain the cells. The events per second and

gating used during analysis was critical in increasing resolution sampling. Cells were first

removed from their tubes and, before use in the BD FACS Canto II, each sample was pipetted

through a 30µm nylon mesh, to remove cell clumping, and was then placed into a FACS sample

tube and run through the FACS Canto II on low speed setting which significantly increased cell

resolution over the medium or high speed sampling. The flow cytometer fluoresced cells at

Page 102: Investigating the effects of, the isoflavone, Phenoxodiol

82

488nm excitation and detected at 619nm emission. Cells were detected and recorded in linear

mode.

Once cell fluorescence was detected the laser voltage settings were adjusted to move cells into

the gated area, voltage was adjusted per run dependent upon cell type and moved cells on the

PI_DNA-A (y-axis) versus PI-A (x-axis) graph towards a 45 degree. Once the voltage was set,

cells were gated to exclude doublets, triplets and cell debris using a SSC-A versus FSC-A graph.

FSC-A and FSC-H population counts were determined by the SSC-A versus FSC-A graph and the

populations were then equalised by altering FSC area scaling until each was displaying a close

(≥95%) population count. FSC was then determined to be accurately configured. Once FSC-A

and FSC-H had been configured PI_DNA-A and PI-DNA-H detection was equalised by gating a

SSC-A versus PI_DNA-A graph and placing a gate over the G1 population and using this

population to determine the PI_DNA-A and PI_DNA-H population, which was configured by

altering blue laser scaling until the population counts were close (≥95%) to each other and it

was determined that cells were accurately configured.

Page 103: Investigating the effects of, the isoflavone, Phenoxodiol

83

FIGURE 13 CELL POPULATIONS MULTI-GATED IN FACS DIVA SOFTWARE FOR ACQUISITION AND

ANALYSIS

Gated populations used to determine FSC and PI_DNA for set-up of the FACS Canto II to detect

DNA height and width in cells accurately.

Page 104: Investigating the effects of, the isoflavone, Phenoxodiol

84

FIGURE 14 CELL CYCLE POPULATIONS FOR ANALYSIS AQUIRED USING GATED POPULATIONS IN FACS

DIVA SOFTWARE

FSC and PI_DNA once they have been adjusted by altering FSC Scaling and Blue Laser scaling.

Following successful setup cells were processed through the FACS Canto II until at least 10,000

events lay within the P3 gate. It is evident from the count versus PI_DNA-A graph that the cells

have left shifted along the axis, this did not alter analysis as cell lines would shift left and right

slightly based on sort rate, volume and would be compensated for in FlowJo.

Page 105: Investigating the effects of, the isoflavone, Phenoxodiol

85

3.8.3. CELL CYCLE DATA ANALYSIS

Analysis of cell cycle results was performed using FlowJo software version 7.2.5, engine

1.999995. Cells were entered into a new workspace per experiment and each sample was

gated first with FSC area (x-axis) and SSC area (y-axis) and then sub-gated looking at DNA Area

(x-axis) and DNA Width (y-axis) before a cell cycle analysis was performed on the DNA sub-

gated populations.

FIGURE 15 EXAMPLE OF CELL POPULATION GATING FOR ANALYSIS IN FLOWJO SOFTWARE

Example of cell population gating indicating how the side scatter/forward scatter population

gate can remove doublet cells, triplet cells and small debris fields while DNA width/DNA area

gate can be used to further remove cellular debris.

Once a cell cycle analysis had been performed a Watson (pragmatic) analysis was performed

with draw model sum, draw components and pattern fill turned on. With G2 populations

assumed to be between 1.95 and 2.05 of the G1 population, dependent upon the cell line, and

a serious change in G2 population and overall cell cycle shape, the Watson (pragmatic) analysis

performed better in evaluating treatment effects upon cell cycle distribution, than the Dean-

Jett-Fox (DJF) method. As an analysis the Watson (pragmatic) makes no assumptions about the

shape of the s-phase distribution and fits the s-phase under the graph while the Dean-Jett-Fox

Page 106: Investigating the effects of, the isoflavone, Phenoxodiol

86

analysis assumes that the s-phase can be modelled by a second degree polynomial, requiring a

G2 peak which was lacking in some treatments. Following analysis with FlowJo results were

tabulated in Microsoft Excel 2007 and graphed using column and pie graph options.

FIGURE 16 WATSON VERSUS DEAN-JETT-FOX CELL CYCLE ANALYSIS IN FLOWJO SOFTWARE

A Watson (pragmatic) versus Dean-Jett-Fox (DJF) analysis of a control cell culture, clearly visible

is the Gaussian distribution of the DJF S-phase as assumed by the G1 and G2 peak distribution

versus the Watson (pragmatic) method of S-phase distribution defined by graph shape.G1

phase populations are represented in green, S phase populations in yellow and G2 phase

populations in blue.

WATSO N (PRAG MATI C) C EL L CYCLE DI STRIBUTIO N

DEAN-JETT - FO X CELL CY CL E DI STRIBUTIO N

G1

S G2

Page 107: Investigating the effects of, the isoflavone, Phenoxodiol

87

FIGURE 17 WATSON PRAGMATIC CELL CYCLE ANALYSIS OF PC3 CELLS UNDERGOING PHENOXODIOL

TREATMENT

The cell cycle of a PC3 cell treated with 30µM Phenoxodiol for 48 hours. The G2 peak is almost

undetectable and only found by constraining samples so that G2 = 1.95-2.05 x G1. Analysis with

DJF fails to determine S and G2 phases successfully, as determining a Gaussian distribution

pattern is impossible without a G2 peak. Therefore DJF analysis cannot be used to on

Phenoxodiol treated samples.

3.9. ASSESSMENT OF RNA EXPRESSION

Analysis of RNA (ribonucleic acid) expression, targeting genes of interest, was performed

through the use of the Polymerase Chain Reaction (PCR) technique. The technique utilised the

Taq polymerase enzyme to make copies of messenger RNA (mRNA) genes, designated by

primer sequences, and to use these copies to determine changes between control and

treatment sample groups. This technique allowed measurement of expression levels of specific

genes of interest which determined whether phenoxodiol was affecting particular signalling

pathways within the cells. To perform this analysis cells would undergo RNA extraction, DNAse

Treatment, Reverse Transcription PCR (RT-PCR), RT-PCR cleanup and finally qPCR testing.

Page 108: Investigating the effects of, the isoflavone, Phenoxodiol

88

3.9.1. RNA EXTRACTION

Cells underwent RNA extraction as the first step in mRNA analysis. Cells were seeded in 6-well

plates, n=4, at appropriate seeding rates (stated above), and grown for 48 hours. Following 48

hours of growth the media was removed and replaced with 2mL of treatment media

containing; Vehicle control (DMSO), 10µM PXD or 30µM PXD. After 24 and 48 hours treatment,

the media was removed, placed in a 2mL eppendorf tube and centrifuged at 800g. The media

was then discarded leaving cell pellet behind while 1mL of Tri Reagent (Molecular Research

Centre Cat#TR118, now Tri Reagent RT Cat#RT111) was placed into each well. Tri Reagent

works via the Chomczynski method of RNA separation using Guanidinium thiocyanate-phenol-

chloroform phase separation of DNA, RNA and protein (Chomczynski and Sacchi 1987). This

method takes longer than a column based separation technique but has been shown to

increase RNA purity.

Wells were scraped with a cell scraper (Sarstedt) and pipetted into the 2mL eppendorf tube

that had their media centrifuged in, then mixed vigorously with a 1000µL pipette (Eppendorf

Cat#3121 000.120). Homogenised cells were rested at room temperature for 5 minutes before

the addition of 200µL chloroform (Sigma-Aldrich Cat#472476) was added to each sample.

Tubes were shaken for 30 seconds before resting at room temperature for 10 minutes. Tubes

were then centrifuged for 15 minutes at 12,000g 4°C with the top layer (typically 70% of the

sample) of the phase separated mixture being transferred to a new tube leaving behind a DNA

and protein solution. 500µL of isopropanol was then added to the top phase that had been

removed from each sample and was vortexed then stored overnight at 4°C.

Following overnight storage samples were centrifuged for 10 minutes at 4°C and 12000g. A

white pellet forms at the base of the tube, the supernatant is removed leaving behind the

Page 109: Investigating the effects of, the isoflavone, Phenoxodiol

89

pellet and 1mL of -20°C 75% Ethanol was added to each tube, samples were vortexed then

centrifuged again for 5 minutes at 4°C and 12,000g. Following the 75% ethanol wash step the

ethanol was removed leaving behind a pellet which was air dried in a class 1 biohazard hood

for approximately 5 minutes before the addition of 50µL of DEPC treated water (ribonuclease

free).

3.9.2. RNA INTEGRITY

RNA sample purity and concentration was determined by taking a 2µL aliquot of each sample

for spectrophotometer analysis on a Nanodrop 1000 spectrophotometer (ThermoScientific

cat#ND1000) with samples outside of a 1.7-2 260/280nm range discarded and those within

stored at -80°C. RNA also underwent electrophoresis on a 1.5% agarose-TAE gel with ethidium

bromide or Sybrsafe (Invitrogen Cat#S33102), as the preferred nucleotide fluorescent stain,

added to the gel before it was set in a mould. Five microliters of RNA was combined with 1µL

of 6x Blue/Orange Loading Dye (Promega Cat#G1881) and placed in the gel under a 1X TAE

buffer at 100 volts for 30 minutes. Visualisation of the 28s, 18s, and 5s ribosomal bands was

established under ultra violet light using a Kodak 2000MM Image Station. Samples were stored

at -80°C. Sample purity and concentration was determined by taking a 2µL aliquot of each

sample for spectrophotometer analysis on a Nanodrop spectrophotometer with samples

outside of a 1.7-1.9 260/280nm range discarded.

3.9.3. DNASE TREATMENT

RNA was DNase treated to remove any genomic DNA before use, with an RQ1 RNase-free

DNase treatment (Promega Cat#M6101). The quantity of sample required for 2µg of RNA was

determined with a spectrophotometer then 2µL of RQ1 DNase and 1µL of RQ1 DNase 10X

Page 110: Investigating the effects of, the isoflavone, Phenoxodiol

90

Reaction buffer15 was added to the RNA in 200µL clear Axygen PCR tubes (Fisher Biotec

Cat#PCR-02-C) and made up to a final volume of 10µL of DEPC treated, nuclease free water.

The solution was vortexed and placed in a PTC-100 thermocycler (MJ Research Inc Cat#PTC-

100 ) at 37°C for 30 minutes. Following DNase treatment 1µL of DNase Stop solution16 was

added to each tube, mixed via vortexing and the tubes returned to the thermal cycler for

another 10 minutes at 65°C to inhibit the activity of the DNase leaving a 2µg solution of RNA

with a volume of 11µL.

TABLE 13 RQ1 DNASE TREATMENT REACTION COMPONENTS

Reagent Volume

2µg RNA __µL

RQ1 RNase Free DNase 2µL

RQ1 10x Reaction Buffer 1µL

H2O Volume to final reaction volume of 10µL

3.9.4. REVERSE TRANSCRIPTASE POLYMERASE CHAIN REACTION

Reverse transcriptase polymerase chain reaction (RT PCR) is a technique which takes purified

RNA and amplifies the low yield RNA whilst also producing a complementary strand referred to

as complementary DNA (cDNA). This strand allows for increased stability of mRNA as well as

the subsequent amplification increasing the ability to perform PCR analysis of expression of

specific mRNA sequences within a sample.

One microgram of DNase treated RNA was processed through RT PCR using a M-MLV Reverse

Transcriptase RNase H Minus, Point Mutant Taq enzyme (Promega Cat#M3682). Briefly, 1g of

15

400mM Tris-HCl (pH 8.0), 100mM MgSO4 and 10mM CaCl2. 16

20mM EGTA (pH 8.0)

Page 111: Investigating the effects of, the isoflavone, Phenoxodiol

91

DNase treated RNA (5.5µL) had 8µL DEPC H2O and 0.5µL Random Primers (Promega Cat#

C1181) added, mixed gently and heated to 70°C for five minutes and incubated on ice for five

minutes. The following components were then added in order to give a total reaction volume

of 25µL.

TABLE 14 RT PCR REACTION COMPONENTS

Reagent Volume

M-MLV 5x Reaction Buffer17 5µL

10mM dNTP’s18 1.3µL

M-MLV RNase H Minus, Point Mutant 1µL

RNasin19 1µL

DEPC H2O 2.7µL

Following the addition of all components into the appropriate PCR tube the PTC-100

thermocycler was programmed to run for 10 minutes at 25°C, 50 minutes at 55°C and finally

for 15 minutes at 70°C before holding at 4°C, samples were then stored at -20°0C.

3.9.5. POST PCR CLEAN-UP

Following RT PCR a clean-up of excess dNTP’s, reaction salts, M-MLV Taq and unused primers

from the sample cDNA was performed using an UltraClean PCR clean-up kit (MoBio

Laboratories, Inc. Cat#12500). Each PCR reaction (25µL) was treated with a 5-times volume

(125µl) of SpinBind (Guanidine-HCl/Isopropanol), mixed well and transferred to a spin filter

unit and centrifuged at 13000rpm for 30 seconds. The flow-through was discarded, the filter

basket returned to the tube and 300µl of SpinClean (<80% Ethanol solution) added to the filter.

17

At final concentration: 50mM Tris-HCl (pH 8.3), 75mM KCl, 3mM MgCl2, 10mM DTT 18

Promega Cat#C1145 19

Promega Cat#N2111

Page 112: Investigating the effects of, the isoflavone, Phenoxodiol

92

The spin column was then spun once for 30 seconds, with flowthrough discarded, then again

for one minute. The filter basket was then removed to a clean 1.5ml tube and 50µl of Elution

Buffer (10mM Tris, pH 8.0) added directly to the filter. The spin column was centrifuged at

13000 rpm for 30 seconds, the filter basket disposed of, with the resultant cDNA stored in a

freezer at -20°C.

3.9.6. PRIMER DESIGN

Before qPCR can occur primer sets must be designed to bind to the cDNA of the genes of

interest exclusively. Primers are base pair sequences of arginine, guanine, cytosine and

tyrosine that are ~20 base pairs long and are designed to be sequence specific to a gene of

interest. Human nucleotide sequences were fetched from the Pubmed database

(http://www.ncbi.nlm.nih.gov/pubmed/) and primers were checked or designed to bind to the

gene sequence, or sequences if there were multiple available. Primers not available from

published sources were designed using the online software (Primer 3,

http://frodo.wi.mit.edu/cgi-bin/primer3/primer3_www.cgi) and rechecked using local

software. All primers underwent a BLAST (http://blast.ncbi.nlm.nih.gov/Blast.cgi) search check

to ensure no cross reactivity. Primers not available from published works were designed to

span at least one intron/exon boundary with a primer sequence specifically crossing the

boundary to prevent the amplification of genomic DNA and ensure only cDNA replication.

Wherever possible the GC content between primer sets was kept as close as possible to allow

for equivalent annealing temperatures while primer designed with large regions of matching

base pairs were excluded to limit primer dimerisation as much as possible.

Page 113: Investigating the effects of, the isoflavone, Phenoxodiol

93

TABLE 15 PRIMER SEQUENCES, PRODUCT SIZE AND ANNEALING TEMPERATURE

Gene Sequence Product Size

(Base Pairs)

Annealing

Temperature

AIF Forward: GATCACGCTGTTGTGAGTGG

Reverse: TCTTGTGCAGTTGCTTTTGC

179bp 61°C

β-Catenin Forward: GATTTGATGGAGTTGGAC

Reverse: TGTTCTTGAGTGAAGGAC

218bp 52°C

Bax Forward: GCTGGACATTGGACTTCCTC

Reverse: TCAGCCCATCTTCTTCCAGA

167bp 61°C

Bcl-xL Forward: ACAATGCAGCAGCCGAGAG

Reverse: ATGTGGTGGAGCAGAGAAGG

167bp 61°C

Caspase 3 Forward:

AAGGATCCTTAATAAAGGTATCCATGGAGAACACT

Reverse:

AAAGAATTCCATCACGCATCAATTCCACAATTTCTT

322bp 55°C

Cyclin-D1 Forward: AACTACCTGGACCGCTTCCT

Reverse: CCACTTGAGCTTGTTCACCA

165bp 62°C

GAPDH Forward: CAGAACATCATCCCTGCATCCACT

Reverse: GTTGCTGTTGAAGTCACAGGAGAC

185bp 60°C

Ki-67 Forward: AGTCAGACCCAGTGGACACC

Reverse: TGCTGCCGGTTAAGTTCTCT

225bp 60°C

L19 Forward: CTGAAGGTCAAAGGGAATGTG

Reverse: GGACAGAGTCTTGATGATCTC

194bp 52°C

p21WAF1 Forward: CCGAAGTCAGTTCCTTGTGG

Reverse: AAGTCGAAGTTCCATCGCTCA

333bp 61°C

sFRP4 Forward: CGATCGGTGCAAGTGTAAAA

Reverse: GACTTGAGTTCGAGGGATGG

181bp 56°C

XIAP Forward: GGGGTTCAGTTTCAAGGACA

Reverse: CGCCTTAGCTGCTCTTCAGT

183bp 56°C

3.9.7. REAL TIME QUANTITATIVE PCR ANALYSIS

Polymerase chain reaction (PCR) is a cycling of temperatures that results in the amplification of

a specific product and shows if a particular gene is transcriptionally expressed. Quantitative

PCR (qPCR) is similar to PCR but has an additional fluorescent tag which binds to double

Page 114: Investigating the effects of, the isoflavone, Phenoxodiol

94

stranded DNA and provides, fluorometrically, a quantitative profile on the amplified product.

Each annealed primer-cDNA set denatures into single strands within a small temperature

range according to its length, sequence and GC content (Ririe et al. 1997).

Real time quantitative PCR (qPCR) analysis was performed on the clean cDNA samples to

detect and determine levels of gene expression within the cell lines, under treatment and

vehicle control conditions, over 24 and 48 hours of treatment. In a regular PCR the mRNA is

mixed with buffer, magnesium, dNTP’s, Taq enzyme and primers then undergoes repeated

cycles of denaturation of the mRNA, annealing of the primers to the denatured mRNA and

extension, which occur at different temperatures. The product is a specifically amplified region

of mRNA.

Quantitative PCR was initially performed upon samples to provide a viable product which was

size confirmed using gel electrophoresis. Following electrophoresis samples were excised and

purified to produce a saturated PCR positive sample which was serially diluted 10 fold in 1xTE

(pH 8.0) until a set of standards had been produced which were run in duplicate in all reactions

allowing accurate quantitation of qPCR product concentration. Corbett Life Science RG3000

and RG6000 (now Qiagen Rotor-Gene Q) real time PCR thermocyclers were used to perform all

qPCR experiments. Quantitative PCR was conducted using two separate Taq enzymes; iQ Sybr

Green Supermix 2X (Biorad, Cat#170-8882) and Immolase Taq (Bioline Cat#BIO-21046).

iQ Sybr Green based qPCR’s were performed with a reaction mix of; 5μl 2X iQ Sybr Green

Supermix20, 1μl forward primer, 1μl reverse primer, 1μl DEPC H2O and 2μl cDNA. Immolase

20

100mM KCl, 40mM Tris-HCI, pH 8.4, 0.4mM of each dNTP (dATP, dCTP, dGTP, dTTP), iTaq DNA polymerase 50 units/ml, 6mM MgCl2, SYBR Green I, 20nM fluoresein, and stabilisers.

Page 115: Investigating the effects of, the isoflavone, Phenoxodiol

95

reactions had a reaction mix as seen in Table 14. Mixtures were denatured at 95°C for 10

minutes before cycling at the following steps; 95°C denature for 2 seconds, annealing at

temperate specified in Table 1321 for 15 seconds and extension at 72°C for 10 seconds. Cycling

was performed for between 40 and 45 total cycles after which a stepwise melt curve was

performed.

TABLE 16 IMMOLASE TAQ QPCR REACTION MIX

Component Volume for 10µL Reaction (µL)

10X Buffer22 1.0

Immolase Taq 0.1

50mM MgCl2 0.5-0.8 Depending on primer set

Sybr Green 0.5

10mM dNTP’s 0.4

Forward Primer 0.4

Reverse Primer 0.4

DEPC H2O 4.7-5.7

cDNA 1.0-2.0

Melt curves were performed for each qPCR reaction by heating the reaction in a 0.5°C

stepwise fashion from 72°C up to 99°C to ensure the exclusive melting of the desired product.

The fluorescence of the sample is plotted against the temperature, with a typical spike in

fluorescence occurring at the temperature at which the product denatures. Accordingly, the

desired product can be identified, as well as any contaminants such as primer dimerisation

(primer-to-primer binding). The melt curve and standards R2 value were used to determine the

quality of the sample run and reactions with high levels of multiple product formation and low

R2 values were not quantified.

21

Page Number 61 22

Proprietary Buffer, reagent mix unavailable.

Page 116: Investigating the effects of, the isoflavone, Phenoxodiol

96

FIGURE 18 EXAMPLE OF A CYCLIN-D1 MELT CURVE

3.9.8. GEL ELECTROPHORESIS AND EXTRACTION

Samples underwent gel electrophoresis against a known 100bp DNA Step Ladder (Promega

Cat#G6951) to confirm product size and further ensure if primer dimerisation did occur at the

same melt temperature as the gene of interest it was detected due to size differences.

Samples underwent electrophoresis on a 1.5% agarose-TAE gel with 5µL per 50mL 10,000x

GelRed (Biotium Cat#41002) nucleotide fluorescent stain added to the gel before it was set in a

mould. Five microliters of sample was combined with 1µL of 6x Blue/Orange Loading Dye

(Promega Cat#G1881) and placed in the gel under a 1x TAE buffer at 100 volts for 30 minutes.

Visualisation of the 100bp DNA Step Ladder bands was established under ultra violet light

using a UV illuminator and the appropriate product band was excised from the gel carefully to

prevent UV irradiation.

Using an Axyprep DNA gel extraction kit (Axygen Biosciences Cat#AP-GX-250) the excised gel

band was placed into a pre weighed 2mL eppendorf tube, and the total weight was calculated

and then removed from tube-only weight to work out gel-only weight. Once the gel weight

was determined the gel extraction kit used an assumption of 100mg = 100µL of reagent

volume. Once weighed, 3x sample volume of buffer DE-A was added to the tube and the

Page 117: Investigating the effects of, the isoflavone, Phenoxodiol

97

gel/buffer solution was heated at 75°C and vortexed until gel was fully solubilised. Once mixed

0.5x DE-A volume of buffer DE-B was added and if the fragment was <400bp then a further 1x

sample volume of isopropanol was added to the tube. An Axyprep spin column was placed into

a new 2mL eppendorf tube and the dissolved gel solution placed into the spin column then

centrifuged at 12,000g for 1 minute. Filtrate was discarded and 500µL of buffer W1 was added

through the spin column, sample was centrifuged at 12,000g for 30 seconds. The filtrate was

discarded again and 700µL of buffer W2 added to the column and centrifuged for a further 30

seconds at 12,000g. The filtrate was again discarded and the column centrifuged for 60

seconds at 12,000g to ensure complete drying of the filter and removal of any residual buffer.

The column was then placed into a new 1.5mL micro-centrifuge tube and 25µL of elutent

pipetted onto the membrane, left to stand for 1 minute then centrifuged at 12,000g for 1

minute. The DNA was stored at -20°C in a post-PCR room for potential sequencing and to be

serially diluted into PCR standards.

3.9.9. QPCR STANDARD PRODUCTION

Standards were made for each gene to enable accurate quantification of products for qPCR

reactions. Tenfold serial dilutions were made from gel extracted DNA standards were then

used simultaneously in qPCRs to establish the relative expression of the cDNA of interest for

each of the primer sets tested. Standards were cycled in duplicate and a standard curve was

created to generate an R2≥0.99, with the experimental samples assigned a value depending on

where they fit on the standard curve. By standardising these results against the human

ribosomal L19 internal control, the amount of product in each sample is quantifiable. Samples

were assigned values via rotor-gene software, Corbett Rotor-Gene 6000 application software,

version 1.7 (Build 87).

Page 118: Investigating the effects of, the isoflavone, Phenoxodiol

98

TABLE 17 QPCR STANDARDS SERIAL DILUTION

Standard Serial Dilution

A Gel extracted sample

B 1µL Standard A + 9µL 1x TE

C 1µL Standard B + 9µL 1x TE

D 1µL Standard C + 9µL 1x TE

E 1µL Standard D + 9µL 1x TE

F 1µL Standard E + 9µL 1x TE

G 1µL Standard F + 9µL 1x TE

H 1µL Standard G + 9µL 1x TE

FIGURE 19 QPCR STANDARDS CYCLING ASSESSMENT

Standards were 10 fold dilutions of each other, those with samples within their bounds were

selected as the most appropriate standards for samples. Cycling analysis of standards allowed

inappropriate replicates to be removed before analysis. Y axis contains concentration while X

axis displays at which cycle detectable amplification of product began.

Page 119: Investigating the effects of, the isoflavone, Phenoxodiol

99

FIGURE 20 STANDARDS ANALYSIS BY CORBETT ROTOR-GENE 6000 SOFTWARE.

Standards analysis from the Corbett rotor-gene software uses the standards to design a line of

best fit with an R2≤.99 and apply this to cycling samples. Y axis contains a given count while X

axis displays listed concentrations set by user.

FIGURE 21 QPCR FULL RUN STANDARDS AND SAMPLES

A qPCR experiment displaying both standards and samples with each other, negative samples

are not displayed as they do not cycle.

Page 120: Investigating the effects of, the isoflavone, Phenoxodiol

100

FIGURE 22 STANDARDS AND SAMPLES ANALYSIS BY CORBETT ROTOR-GENE 6000 SOFTWARE

A completed run with samples and standards displayed against the generated line of best fit.

3.10. ASSESSMENT OF PROTEIN EXPRESSION

Protein expression changes following treatment with Phenoxodiol was determined by analysis

using the western blot technique. Protein analysis by Western blot shows whether a particular

protein is translationally expressed for the epitote that is specific to the particular antibody

used. Western blot analysis was performed using polyacrylamide gel electrophoresis (PAGE) to

separate different size proteins which were then transferred to a nitrocellulose membrane

where primary antibodies were bound to the epitope of proteins followed by a secondary

tagged antibody to visualise proteins of interest. The cytoskeleton protein β-actin was used as

a control standard to prevent variation in results due to loading or concentration differences.

Page 121: Investigating the effects of, the isoflavone, Phenoxodiol

101

3.10.1. PROTEIN EXTRACTION

Protein was extracted from all cell lines using a Radioimmunoprecipitation (RIPA) Buffer

consisting of 150mM NaCl, 50mM Tris-HCl (pH 7.5), 1% Triton X-100, 0.5% Sodium

Deoxycholate, 0.1% SDS and 0.1mM Phenylmethylsulfonyl fluoride (PMSF) in dH2O. Cells were

seeded into either 6 well, or 10cm dishes, at appropriate concentrations then covered with

2mL of complete cell media and incubated at 37°C 5% CO2 for 48 hours. Media was then

aspirated off and replaced with appropriate volumes of treatment media (2mL per well, in a 6

well plate, or 5mL per 10cm dishes) and incubated for either 24 or 48 hours at 37°C 5% CO2.

RIPA buffer was prepared directly before addition to the wells containing cells for every

extraction by addition in order of; 50mg sodium deoxycholate, 1.5ml 1M NaCl, 100μl of 10%

SDS, 500μl 1M Tris-HCl (pH 7.5) and 100μl Triton X-100 dissolved in 7.8ml of dH2O. Following

preparation buffer was placed into -20°C to chill to as cold as possible without freezing before

use, immediately prior to use 10μl of -20°C 100mM PMSF was added.

To prepare samples for use media was aspirated off the 6-well/10cm dishes and placed into a

2mL/15mL eppendorf tube and centrifuged for 5 minutes at 800g. The media in the tubes was

then aspirated post centrifuge leaving the cell pellet behind. Either 1mL or 3mL of sterile 4°C

1xPBS (pH 7.4) buffer was added to each well/dish and then the well/dish was scraped with a

cell scraper and resulting cell and 1xPBS mixture pipetted into the 2mL/15mL eppendorf tubes

where appropriate, then centrifuged for 5 minutes at 800g. The addition of the 1xPBS and

scrape steps was critical for removing excess FBS from samples before RIPA buffer was added,

as FBS was found to interfere with accurate protein extraction causing the western blot

analysis to fail. Once the cell pellet was in the tube and the supernatant removed, 50µL RIPA

buffer was added to 6-well plate samples or 200µL RIPA buffer added to 10cm dish samples, all

were mixed and cells lysed via vigorous pipetting and vortexing before being transferred to

new 1.5mL eppendorf tubes. Samples were then left to sit on ice at for 30 minutes before

Page 122: Investigating the effects of, the isoflavone, Phenoxodiol

102

being placed into a 4°C centrifuge and centrifuged at maximum speed for 5 minutes.

Supernatant was transferred to a new tube and stored at -80°C for use in protein

quantification and western blot analysis.

3.10.2. BRADFORD PROTEIN QUANTIFICATION ASSAY

The Bradford protein quantification assay uses a standard curve to help determine

concentration of protein in samples. Protein standards were made by serial dilution of

acetylated Bovine Serum Albumin (BSA 10mg/ml) to create standards ranging from 500 to

100μg/ml protein.

TABLE 18 BRADFORD PROTEIN STANDARDS SETUP

Solution Concentration of BSA Components

A 500µg/mL 50µL BSA (10mg/mL) + 950µL

0.01xPBS

B 400µg/mL 760µL A + 190µL 0.01xPBS

C 300µg/mL 750µL B + 250µL 0.01xPBS

D 200µg/mL 600µL C + 300µL 0.01xPBS

E 100µg/mL 400µL D + 400µL 0.01xPBS

Protein samples were diluted 1:20 in 0.01M PBS and 10μl transferred into glass tubes. In

duplicate, 10μl of each Standard was transferred to glass tubes, to which 200μl of protein dye

was added, vortexed and incubate for 5 minutes at room temperature. The colour of the

protein samples was checked against the Standards to ensure their colour would fit within the

standard curve. 0.01M PBS was used to zero the nanodrop spectrophotometer. Standards

were analysed by 2µL addition to the spectrophotometer and using a visible light analysis a

standard curve produced. Sample concentration was determined by adding 2μl to the

Page 123: Investigating the effects of, the isoflavone, Phenoxodiol

103

nanodrop and analysing with the spectrophotometer, and the concentration of protein (μg/ml)

was extrapolated from the standard curve and then multiplied by 20 to remove the initial

dilution factor.

3.10.3. SDS-PAGE WESTERN BLOT ANALYSIS

Protein analysis was performed using a western blot analysis whereby protein expression is

determined through antibody binding to specific epitopes on target proteins followed by

chemo-luminescent visualisation of the antibody binding. Once visualised, quantification of

samples can occur by the comparative luminosity of treatment versus control accounting for

the β-actin loading control. Sodium Dodecyl Sulfate Polyacrylamide Gel Electrophoresis (SDS-

PAGE) is the process where protein samples are separated into fragments according to their

molecular weight, this occurs by loading protein into an acrylamide gel and running a current

through the gel. Proteins move through the gel at a speed designated by their molecular

weight, charge and conformation. A Biorad Mini Trans-Blot Electrophoretic Transfer Cell

apparatus (Bio-Rad Cat#170-3930) was used to perform both SDS-PAGE protein separation and

the nitrocellulose membrane transfer.

Glass plates with 1mm spacers, their accompanying smaller glass plates and the 12 well combs

were cleaned with 70% ethanol, allowed to dry then aligned and placed into the stacking gel

locking mechanism. A 12% acrylamide separating gel was poured into the space between the

two glass plates until separating gel was approximately 1.5cm below the top of the glass

plates, 200µL of dH2O was then added to each corner of the separating gel and left for 20

minutes to set. Blotting paper was then used to remove excess dH2O and unset gel, if the

separating gel was not completely flat across the top surface the gel was destroyed, glass

plates cleaned and another separating gel poured. Once the separating gel was correctly

Page 124: Investigating the effects of, the isoflavone, Phenoxodiol

104

formed, a 4% acrylamide stacking gel was poured across the top of the separating gel, with the

12 well comb placed into it, ensuring no air bubbles formed at the base of the comb, to

prevent issues with protein loading.

TABLE 19 SDS-PAGE/ACRYLAMIDE GEL RECIPE

Gel Ingredient Volume, 12%

Separating Gel

Volume, 4% Stacking

Gel

ddH2O 3.3mL 6.1mL

1.5M Tris-HCl (pH 8.8) 2.5mL -

0.5M Tris-HCl (pH 8.8) - 2.5mL

10% SDS 100µL 100µL

30% Acrylamide/Bis23 4.0mL 1.33mL

10% Ammonium Persulfate (APS) 100µL 50µL

Tetramethylethylenediamine

(TEMED)24

10µL 10µL

During separating and stacking gel solution mixing the; dH2O, 1.5M/0.5M Tris-HCl (pH 8.8),

30% Acrylamide/Bis, 10% SDS was added and mixed together and just before the pouring of

the gel solution the N,N,N',N'-Tetramethylethylenediamine (TEMED) and 10% ammonium

persulfate (APS) was added then quickly poured as TEMED and APS cause the gel to set. The

separating and stacking gel solutions were mixed in 15mL falcon tubes and discarded in

incineration containers once set or unused as acrylamide is a neurotoxin and cannot be

discarded safely in another manner. The stacking gel was allowed to set for 20 minutes then

glass plates were removed from gel locking mechanism and placed into the electrophoresis

tank in a new gel electrophoresis locking mechanism. Gels were placed into the

electrophoresis tank with the smaller plates facing each other, and the reservoir between

23

Cat# A3574 – Sigma Aldrich 24

Cat# T9281 – Sigma Aldrich

Page 125: Investigating the effects of, the isoflavone, Phenoxodiol

105

them filled with 4°C electrode buffer25. Electrode Buffer was also added to the electrophoresis

tank to a final depth of 3cm to cover the electrodes within the bottom of the tank and placed

between the two glass plates creating a layer of buffer over the comb/well level.

Three separate samples per treatment group (nine samples total per time point, 18 total

samples per cell line) were thawed and 30 or 50µg of each sample was placed into new 1.5mL

eppendorf tubes. The most dilute sample equalled the largest volume and all other tubes had

0.01M PBS added up to equal the volume. Once the volume was determined an equal volume

of protein loading buffer26 was added to each sample and then gently mixed before being

placed in a 95°C heat block for a maximum of 5 minutes, to denature the protein. Samples

were removed from the heat block and loaded, with a Hamilton microlitre syringe, into

consecutive wells in the set acrylamide stacking gel.

TABLE 20 EXAMPLE OF PROTEIN SAMPLE PREPARATION

Sample Volume for 30µg 0.01M PBS Loading Buffer26

Control 1 14µL 6µL 20µL

Control 2 18µL 2µL 20µL

Control 3 12µL 8µL 20µL

PXD 10µM 1 17µL 3µL 20µL

PXD 10µM 2 16µL 4µL 20µL

PXD 10µM 3 19µL 1µL 20µL

PXD 30µM 1 20µL - 20µL

PXD 30µM 2 12µL 8µL 20µL

PXD 30µM 3 14µL 6µL 20µL

25

25mM Tris-HCl, 200mM Glycine and 0.1% SDS 26

4% SDS, 2% β-mercaptoethanol, 20% v/v glycerol, 250mM Tris-HCl (pH 6.8) and 0.006% bromophenolblue

Page 126: Investigating the effects of, the isoflavone, Phenoxodiol

106

The samples were loaded, underneath the electrode buffer25, into the stacking gel from rows

3-11 with row two containing 5µL of SeeBlue Plus 2 molecular weight standard (Invitrogen,

Cat# LC5925). After loading the protein samples and ladder, the electrodes were inserted into

the apparatus and the electrical source, with the black electrode connected to the negative

terminal and positive electrode connected to the red terminal. The apparatus was set to a

voltage of 100V until the samples reached the top of the separating gel and had started to

cross between stacking and separating gel. The voltage was then increased to 130V and left

until the 6kDa molecular weight band reached the bottom of the separating gel. Once

separated, the gels were removed from the electrophoresis tank and locking mechanism, the

glass plates were gently prised apart to prevent tearing of the gel. Once apart the stacking gel

was removed from the separating gel and the stacking gel was discarded into an incineration

container.

In preparation for the transfer of proteins to Hybond-C Extra nitrocellulose membrane

(Amersham Cat#RPN303E), 4 pieces of blotting paper and 2 nitrocellulose membranes (one for

each gel/timepoint) were cut to the required size (equal to the separating gel) and, with 6

absorbent pads, soaked in transfer buffer27. The transfer cassettes were layered with; one pad,

one piece of blotting paper, the separating gel, the nitrocellulose membrane, more blotting

paper and 2 more pads, with each component layered consecutively upon the black side of the

rack. The contents were kept saturated with transfer buffer throughout and any air bubbles

were removed from between the gel and membrane with a glass rolling pin. The transfer racks

were closed, locked and inserted into a transfer apparatus with the black side facing the back

of the apparatus. A 2-3cm magnetic stirrer was placed into the bottom of the electrophoresis

tank with the transfer apparatus and an -20°C ice brick placed in the tank. The tank was then

27

800ml dH2O, 3.03g Tris, 14.4g Glycine, 200ml Methanol

Page 127: Investigating the effects of, the isoflavone, Phenoxodiol

107

filled with transfer buffer27 and the electrodes attached negative to black, positive to red and

the powerpack set to produce 100V for 75 minutes.

Following the transfer the gel was discarded into an incineration container, and the

nitrocellulose membrane rinsed quickly in Tris-buffered Saline Tween-2028 solution (TBS-T)

prior to being covered with Ponceau S29 solution for 30 seconds, to reveal all transferred

proteins and ensure no air bubble affected areas. The Ponceau S was returned to the stock

bottle, and the excess removed by repeated washes of dH2O. The membrane was cut into

strips containing the desired range of protein sizes for immunoblotting. The remaining

Ponceau S was removed from the membrane with TBS-T for 5 minutes, and the membranes

stored in sealed plastic bags at 4°C until required for immunoblotting.

3.10.4. IMMUNOBLOTTING

In order to prevent non-specific antigen binding, membranes were blocked in 5% non-fat milk

powder (Diploma) in TBS-T for one hour at room temperature. Following blocking membranes

were washed three times for five minutes in TBS-T before 2mL of appropriate primary

antibody (Table 21) was added to membrane. Membranes with primary antibodies were

incubated at 4°C overnight with gentle agitation after which membranes were washed with

TBS-T three times for 5 minutes each to remove any unbound antibody. Streptavidin horse

radish peroxidase (HRP) conjugated secondary antibodies were primarily used. Following

primary incubation membranes were washed in TBST and incubated with; XIAP, β-actin Anti-

Mouse HRP (1:10000; Dako), AIF, Bax, Active β-catenin and sFRP4 Anti-Rabbit HRP (1:10000

Dako), Bcl-xL (1:10000 Vector BA-1000) in TBST for one hour at room temperature. After

28

Tris Buffered Saline/0.05% Tween-20 pH 29

2% Ponceau S, 2%Trichloroacetic acid

Page 128: Investigating the effects of, the isoflavone, Phenoxodiol

108

incubation, excess liquid was removed completely with blotting paper, and 1ml each of the

two SuperSignal West Pico Chemiluminescent Substrates (ECL) (Pierce, Cat#34080), were

combined and placed onto the membrane and incubated for 5 minutes. Excess liquid was

again removed completely with blotting paper, and the membrane bagged and imaged on an

ImageStation (Kodak, IS2000MM).

TABLE 21 ANTIBODY CONCENTRATIONS

Protein 1° Antibody Concentration 2° Antibody Concentration

AIF 1:1000 (Cayman Cat#160773) 1:10000 Anti-Rabbit HRP

(Dako)

BAX 1:200 (Abcam Cat#7977) 1:10000 Anti-Rabbit HRP

(Dako)

Β-actin 1:5000(Sigma Aldrich

Cat#A5441)

1:10000 Anti-Mouse HRP

(Dako)

Β-catenin 1:1000 (Millipore Cat#05-601) 1:10000 Anti-Rabbit HRP

(Dako)

Bcl-xL 1:200 (Abcam Cat#7974) 1:10000 Anti-Rabbit

Biotinylated (Vector BA-1000)

sFRP4 1:1000 (Millipore Cat#09-129) 1:10000 Anti Rabbit (Dako)

xIAP 1:250 (BD Transduction

Cat#610762)

1:10000 Anti-Mouse HRP

(Dako)

Page 129: Investigating the effects of, the isoflavone, Phenoxodiol

109

TABLE 22 IMMUNOBLOTTING PROTOCOL

Protein AIF Bax β-

actin

β-

Catenin

Bcl-xL sFRP4 xIAP

Blocking 5% non-fat milk powder in TBS-T

One hour room temperature with gentle agitation

Antibod

y

Rabbit anti

AIF

(Cayman

Cat#16077

3)

1:1000

TBS-T

Rabbit

anti Bax

(Abcam

Cat#7977

)

1:200

TBS-T

Mouse

anti β-

actin

(Sigma-

Aldrich

Cat#A54

41)

1:5000

TBS-T

Mouse

anti-

active-

β-

catenin

(Upstate

Cat#

1:1000

3% Milk

TBS-T

Rabbit

anti Bcl-xL

(Abcam

Cat#7974

)

1:200

TBS-T

Rabbit

anti

sFRP4

(Millipore

Cat#09-

129)

1:1000

TBS-T

Mouse

Anti xIAP

(BD

Transduc

tion

Cat#6107

62)

1:250 3%

Milk TBS-

T

Overnight at 4°C with gentle agitation

Washing TBS-T

3 x 5 Minutes with gentle agitation

Antibod

y

1:10000

Anti-

Rabbit

HRP

(Dako)

1:10000

Anti-

Rabbit

HRP

(Dako)

1:10000

Anti-

Mouse

HRP

(Dako)

1:10000

Anti-

Mouse

HRP

(Dako)

Anti-

Rabbit

Biotinylat

ed

(Vector

BA-1000)

1:10000

Anti-

Rabbit

HRP

(Dako)

1:10000

Anti-

Mouse

HRP

(Dako)

One hour at Room Temp with gentle agitation

Washing TBS-T

3 x 5 Minutes with gentle agitation

3.10.5. QUANTIFICATION OF WESTERN BLOT ANALYSIS

Western blot protein identification was quantified using the computer program Scion Image

Beta, edition 3B and Kodak 1D Image Analysis Software. The images captured were analysed

Page 130: Investigating the effects of, the isoflavone, Phenoxodiol

110

using the densitometric analysis method where pixel densities are established and

quantitated. Pixel densities of the bands were normalised against β-Actin expression.

3.11. STATISTICAL ANALYSIS

Statistical significance for all experiments was determined using 2-tailed t-tests assuming

unequal variance using Microsoft Excel 2010, data was compared within time points and not

across time points or across cell lines. Statistical significance established at P≤0.05.

Page 131: Investigating the effects of, the isoflavone, Phenoxodiol

111

4 . C H A P TER F O UR: T H E C YTO T OX IC EF F EC TS O F P H ENOX OD I OL ON TH E P ROS TA TE C A NC ER C EL L L INES ; L NC A P , D U1 4 5 A ND PC 3

4.1. INTRODUCTION

Adenocarcinoma of the prostate is the second most commonly diagnosed malignancy in men

and is a common cause of cancer mortality in men in many western countries and increasingly

in developing nations (Hsing et al. 2000). In Australia prostate cancer represents the most

significant of all cancers affecting males with a total of 19,403 new incidences in 2007, 31.3%

of all new cancer cases (AIHW 2010). Prostate cancer also represents the second highest rate

of mortality associated with cancer affecting males at 13% behind lung cancer, which had 19%

of all mortality in 2007 (AIHW 2010). Predictions for the future indicate that prostate cancer

rates are projected to increase from 11,191 in 2001 to 16,800 in 2011, with this large expected

increase reflected in the projection for all cancers affecting males (AIHW 2010). Like other

cancers, prostate cancer has numerous clinical states ranging from a hormone-naïve clinically

localised primary tumour to lethal androgen-independent metastases. Regulation of prostate

growth is mediated via androgens and the corresponding androgen receptor (AR) which

regulates the transcription of target survival and apoptosis genes. Late-phase metastatic

disease is often androgen independent arising from; increased AR expression, enhanced

nuclear localisation of the AR, loss of function due to acquisitions in AR mutation resulting in a

more promiscuous receptor, or the presence of alternative survival pathways (Bcl-xL, Wnt

family upregulation) that circumvent the need for a functional androgen receptor (Gleave et

al. 2005).

Page 132: Investigating the effects of, the isoflavone, Phenoxodiol

112

The primary systemic treatment option for early and late stage prostate cancer is androgen

ablation therapy combined with an apoptosis inducing drug, such as Docetaxel or Paclitaxel.

Chemotherapy treatment options for patients with late-phase metastatic prostate carcinoma

rely on the premise that androgen-insensitive prostate carcinoma cells retain their basic

cellular apoptotic machinery to undergo programmed cell death, however increased tumour-

cell resistance to apoptosis is an underlying molecular reason contributing to disease

progression and chemo-resistance (Miyamoto et al. 2004; Miyamoto et al. 2005). The exact

mechanisms responsible for prostate cancer growth, especially emergence of the androgen-

independent phenotype, are still far from fully understood (Feldman and Feldman 2001).

While other therapies, such as radiotherapy and chemotherapy, are available for advanced

prostate cancer, whether these therapies, either alone or combined with hormonal therapy,

significantly prolong patient survival remains controversial (Miyamoto et al. 2004; Miyamoto

et al. 2005). Thus, novel treatment strategies, instead of, or in combination with, androgen

deprivation therapy for advanced prostate cancer need to be developed such as (Petrylak et al.

2004) (Tannock et al. 2004) have shown.

One such novel method currently undergoing intense investigation is the flavanoid compounds

found in the dietary intake of populations with low cancer rates. Flavanoids are important

regulators in plants of biochemical and physiological processes, acting as antioxidants, enzyme

inhibitors, pigments and hormones. Human consumption of flavanoids has long been

recognised to manage anti-inflammatory, antioxidant, antiallergic, hepatoprotective,

antithrombotic, antiviral and anticarcinogenic activities (Middleton et al. 2000).

Epidemiological studies have consistently shown an inverse association between isoflavone

intake and risk of cancer (Brusselmans et al. 2005). In vitro mechanistic studies on isoflavones

provide insight into modes of anticancer action ranging from cell cycle arrest and apoptosis

induction to angiogenic and antiproliferative effects (Li et al. 2008; Seo et al. 2011). To date,

Page 133: Investigating the effects of, the isoflavone, Phenoxodiol

113

genistein has been shown to demonstrate broad anticancer effects such as receptor tyrosine

kinase and cyclin dependent kinase inhibition (Alhasan et al. 1999; Li and Sarkar 2002).

Phenoxodiol, [2H-1-Benzopyran-7-0,1,3-(4-hydroxyphenyl)], is an isoflavone derivative that has

also been shown to elicit cytotoxic effects against a broad range of human cancers. Currently

undergoing human clinical trials, it has shown promise in patients with recurrent ovarian

cancer where the cancer is refractory or resistant to standard chemotherapy, and in patients

with hormone-refractory prostate cancer (Sapi et al. 2004; Brown and Attardi 2005).

Preliminary studies involving a number of flavanoid derivatives have demonstrated that

Phenoxodiol inhibits cell proliferation of a wide range of human cancer cell lines including,

leukaemia, breast and prostate carcinomas, and is five to twenty times more potent than a

similar compound, Genistein (Aguero et al. 2005). Phenoxodiol has been characterised in

ovarian cancer as affecting key ovarian anti-apoptotic signalling pathways (Kamsteeg et al.

2003) as well as reversing the ability of cells to become resistant to Docetaxel, through over

expression of anti-apoptotic molecules (Sapi et al. 2004). In mammary carcinogenesis

(Constantinou et al. 2003) found that the in vitro activity acted as an inhibitor of cell division

and/or an inducer of cell differentiation whilst short term investigation into Phenoxodiol

activity in prostate cancer by (Axanova et al. 2005) indicated that growth of LNCaP cells, in

monoculture and coculture with osteoblasts, resulted in the downregulation of the cancer

specific enzyme tNADH-Oxidase. Finally (Aguero et al. 2005; Aguero et al. 2010) indicated that

Phenoxodiol induces G1 specific arrest through loss of Cyclin-Dependant Kinase 2 activity by

p53-independent induction of p21WAF1 in a battery of human cell lines, or that phenoxodiol

effects multiple cancer types through prevention of cells from reproducing. Clearly,

phenoxodiol has demonstrated an ability to increase susceptibility of various cancer cell types

to initiate cell cytotoxicity; however the underlying mechanism of action in prostate cancer has

yet to be determined.

Page 134: Investigating the effects of, the isoflavone, Phenoxodiol

114

In this study we investigate Phenoxodiol’s ability to elicit anticancer effects in cells

representative of the clinical stages of prostate cancer development, by directly inhibiting

proliferation, and by eliciting direct cytotoxic effects against androgen-responsive and

androgen-refractory prostate cancer cell lines. We also seek to show that if phenoxodiol elicits

potent cytotoxicity, is it an apoptotic or necrotic response, and is it initiated by classic

apoptotic signalling pathways.

4.2. AIMS

The aims of this chapter were as follows:

Aim 1: To characterise the cell growth rates of cell lines LNCaP, DU145 and PC3 in vitro.

Aim 2: To determine the potential cytotoxicity of phenoxodiol on in vitro prostate cancer cells.

Aim 3: To determine an effective set of doses of phenoxodiol that inflict cytotoxicity on

prostate cancer cells.

Aim 4: To determine the type of cytotoxic cell response following Phenoxodiol treatment in

prostate cancer cells.

4.3. METHODOLOGY

The methodology utilised in this chapter is discussed in detail in the materials and methods

chapter (page 50). The goal of this set of experiments was to determine the optimum growth

characteristics for the three prostate cancer cell lines; LNCaP, DU145 and PC3. The cell lines

are representative of early stage androgen receptor wild type (LNCaP), late stage androgen

receptor mutant (DU145) and late stage androgen receptor null (PC3) prostate cancer.

Page 135: Investigating the effects of, the isoflavone, Phenoxodiol

115

Proliferation assays ensured that after 96 hours of culture the cell lines were undergoing

logarithmic cell growth. Cells were subcultured and seeded onto 96 well plates at varying rates

of initial concentration, after an initial 48 hours of subculture the cultures media were

replaced and a further 48 hours of growth allowed before application of 20µL CellTiter 96®

AQueous One MTS dye. After 3 hours of incubation the colour change was analysed on a

Labskan plate reader at 492nm. Increased absorbance rates were consistent with increased

metabolic and proliferation rates. Once initial seeding concentrations ensuring logarithmic

growth rates over 96 hours were established Phenoxodiol was diluted down into a final

concentration of 10µM and 30µM with a DMSO vehicle. Control and 10µM phenoxodiol

solutions had DMSO added to standardise final vehicle concentration with the 30µM

phenoxodiol treatment. Following dilution the cell proliferation rate, under the influence of

phenoxodiol, was measured by seeding 96 well plates at appropriate rates as pre-determined.

After 48 hours of initial culture the media was replaced for either; DMSO vehicle, 10µM or

30µM phenoxodiol solutions. Solutions were placed in wells lacking cells so that any

interaction of phenoxodiol with MTS dye could be assessed. The cells were treated for 24 and

48 hours, after which MTS dye was added to each well, incubated for 3 hours then assessed for

absorbance.

Acidity analysis was performed on cell media after 48 hours of treatment via a pH balanced

probe. Media pH was measured at 37°C, the incubator temperature for the cells in media.

Functional apoptosis assays to detect early and late stage apoptosis in response to

phenoxodiol treatment were performed. The JC-1 mitochondrial depolarisation assay was used

to determine if phenoxodiol induced mitochondrial efflux, an indicator of early apoptosis. Cells

were seeded into white fluorescent 96 well plates and following phenoxodiol treatment the

media was removed and JC-1 fluorescent dye added with positive control cells receiving a

combination of JC-1 and FCCP. After incubation then 20 minutes of 5% BSA solution the dye

Page 136: Investigating the effects of, the isoflavone, Phenoxodiol

116

was replaced with 100µL PBS and the fluorescent plates analysed on a Fluostar Optima plate

reader. Excitation/Emission was measured from above the plate. Green fluorescence was

measured at 485nm excitation and 520nm emission while red fluorescence was measured at

544nm excitation and 590nm emission. Red emission results were divided by green and the

ratio of red to green compared with a low ratio indicated high depolarisation rate and early

apoptotic induction.

The fluorescent Caspase-3 activity assay measured activated Caspase-3 expression and was

performed by seeding and treating cells on black 96 well fluorescent plates. Following

appropriate treatment with phenoxodiol, activated Caspase-3 was determined using an

EnzChek Capase-3 Assay Kit #2® containing a fluorescent compound (Z-DEVD-R110) which is

converted, by activated Caspase-3, into the green fluorescent rhodamine 110. After treatment

the media replaced with a staining solution and incubated on ice for 30 minutes followed by

room temperature for a final 30 minutes. Excitation/Emission was measured from above the

plate. Cells were measured at 496nm excitation and 520nm emission with a set gain level

between cell lines and time points used.

DNA laddering indicative of late stage apoptosis was performed using the 3’-end labelling

assay. DNA was extracted using a phenol/chloroform extraction protocol and one microgram

of DNA was tagged with radioactive P32 then loaded into 2% agarose gel with loading dye, a

low voltage was applied for approximately 4 hours after which the separated DNA gel was

dehydrated. The gel was exposed to film for 24 hours at -80°C and the film developed. The

dehydrated gel was cut into strips representing the lanes of the gel, loading regions were

removed, and liquid scintillant placed in each gel, after which radioactivity was measured using

a liquid scintillation counter. The exposed film provided a qualitative measure of apoptosis

Page 137: Investigating the effects of, the isoflavone, Phenoxodiol

117

through visible DNA laddering while the scintillation counter provided a quantitative measure

of apoptosis with increased radioactivity correlated to increased apoptotic rate.

An Annexin-V-Fluos/Propidium Iodide double stain apoptosis measurement was performed on

live cells which were placed into a FACSCalibur flow cytometer measuring individual cell

fluorescence. AV/PI double stains indicate normal cell function, early apoptosis and late

apoptosis/necrosis. Cells were grown in 6 well plates and treated with phenoxodiol, after

appropriate time the media was removed and centrifuged. Cells were trypsinised off the 6 well

plates, centrifuged in media then washed with warm PBS before AV/PI staining solution was

added. After 15 minutes incubation at 25°C, running solution was added and the cells analysed

on the FACSCalibur flow cytometer.

Page 138: Investigating the effects of, the isoflavone, Phenoxodiol

118

4.4. RESULTS

Cell proliferation measurements provide an in vitro method of accurately determining cell

growth rates. Figure 23 demonstrates the MTS cell proliferation graphs for prostate cancer cell

lines; LNCaP, DU145 and PC3 over 6, 24 and 48 hours. Cell proliferation was measured at

492nm, analysing the total absorbance of light under the influence of MTS dye, an indicator of

cell metabolism. At 48 hours a visible drop in absorbance represents decreased proliferation in

all cell lines, determining their maximum seed rates. Ninety six well plate seeding rates were

set at; LNCaP: 3000 cells/100µL, DU145: 1500 cells/100µL and PC3: 2500 cells/100µL. This

ensured that after a maximum of 48 hours of treatment (96 hours of total growth) any

decreased proliferative effect that phenoxodiol may have had was not due to a potential

confluence of cells within the 96 well plate confines. After this point all control solutions were

assumed to be at maximum cell proliferation and samples were compared to the control

relative value. Scaling these results 1.8 fold utilised these initial seeding rates for 6 well plate

use; LNCaP; 54,000 cells/mL, DU145: 27,000 cells/mL and PC3: 45,000 cells/mL. When using

10cm plates an initial seeding rate of 100,000 cells/mL was used for all three cell lines, while

not maximising all available space it ensured a T75 flask could be split into more than two

10cm plates.

Page 139: Investigating the effects of, the isoflavone, Phenoxodiol

119

FIGURE 23 CELL PROLIFERATION RATES AT 6, 24 AND 48 HOURS

* indicates initial cell seeding rate with logarithmic cell growth after 48 hours.

0

0.2

0.4

0.6

0.8

1

1.2

1.4

500 1000 1500 2000 2500 3000 4000 5000

Ab

sorb

ance

(4

92

nm

)

Cells Seeded /100uL

LNCaP Cell Proliferation

6hr

24hr

48hr

0

0.5

1

1.5

2

2.5

300 500 1000 1500 2000 2500 3000

Ab

sorb

ance

(4

92

nm

)

Cells Seeded /100uL

DU145 Cell Proliferation

6 hrs

24hrs

48hrs

0

0.5

1

1.5

2

2.5

3

500 1000 1500 2000 2500 3000 4000 5000

Ab

sorb

ance

(4

92

nm

)

Cells Seeded /100uL

PC3 Cell Proliferation

6hr

24hr

48hr

Page 140: Investigating the effects of, the isoflavone, Phenoxodiol

120

Studies have revealed phenoxodiol’s ability to reduce cell proliferation and induce cell

cytotoxicity in cell lines representing various cancer types and mouse xenograft models

(Kamsteeg et al. 2003; Axanova et al. 2005; Aguero et al. 2010). Figure 24 demonstrates the

reduction in prostate cancer cell line proliferation after 24 and 48 hours of phenoxodiol

treatment. Phenoxodiol treatment, in media alone, was found to interact with the MTS dye

and increase the opacity in a dose dependent manner causing a falsely positive increase in cell

metabolism measurements which was accounted for with media only wells. LNCaP cell

treatment with 30µM phenoxodiol significantly decreased cell proliferation, versus control

population, over 24 (p=0.037) and 48 (p=0.01) hours. DU145 cells were significantly reduced in

proliferation rate, versus control population, after 10µM and 30µM treatment concentrations

over 24 (p<0.05, p=0.0032) and 48 (p<0.001, p<0.001) hours of treatment. PC3 cell treatment

with 30µM phenoxodiol significantly decreased cell proliferation, versus control population,

after 24 (p=0.022) hours and both 10µM and 30µM phenoxodiol concentrations after 48

(p<0.05, p=0.002) hours. The resulting proliferation graphs indicated the LNCaP cells were the

least responsive to phenoxodiol treatment with DU145 then PC3 as the most response cell

lines. This result indicated that phenoxodiol was better at arresting cell metabolism in late

stage representative prostate cancer cells (DU145, PC3) than early (LNCaP), utilising the MTS

proliferation assay.

Page 141: Investigating the effects of, the isoflavone, Phenoxodiol

121

FIGURE 24 CELL PROLIFERATION RATES AFTER 24 AND 48 HOURS OF PHENOXODIOL TREATMENT

* indicates significance relative to control for that time point. All cell populations had two

media only replicates of which the mean absorbance value was removed from the total cell

absorbance.

0

20

40

60

80

100

120

24hrs 48hrs

% G

row

th R

ate

Re

lati

ve t

o C

on

tro

l

Treatment

LNCaP Proliferation & Phenoxodiol Treatment

Control

10µM PXD

30µM PXD

0

20

40

60

80

100

120

24hrs 48hrs

% G

row

th R

ate

Re

lati

ve t

o C

on

tro

l

Treatment

DU145 Proliferation & Phenoxodiol Treatment

Control

10µM PXD

30µM PXD

0

20

40

60

80

100

120

24hrs 48hrs

% G

row

th R

ate

Re

lati

ve t

o C

on

tro

l

Treatment

PC3 Proliferation & Phenoxodiol Treatment

Control

10µM PXD

30µM PXD

Page 142: Investigating the effects of, the isoflavone, Phenoxodiol

122

To ensure phenoxodiol was not inducing a reduction in proliferation by acidification or

alkalisation of media the pH of cell culture media was tested. Figure 25 demonstrates the pH

changes induced by 24 and 48 hours of phenoxodiol treatment. No significant differences in

pH were detected after treatment indicating phenoxodiol did not induce cytotoxicity by

acidifying or alkalising the media.

FIGURE 25 PH CHANGES IN PHENOXODIOL TREATED CULTURE MEDIA

0

1

2

3

4

5

6

7

8

24 Hours 48 Hours

pH

Time

LNCaP Media pH Analysis

Control

10µM PXD

30µM PXD

0

1

2

3

4

5

6

7

8

24 Hours 48 Hours

pH

Time

DU145 Media pH Analysis

Control

10µM PXD

30µM PXD

0

1

2

3

4

5

6

7

8

24 Hours 48 Hours

pH

Time

PC3 Media pH Analysis

Control

10µM PXD

30µM PXD

Page 143: Investigating the effects of, the isoflavone, Phenoxodiol

123

The JC-1 mitochondrial depolarisation assay measures induced mitochondrial efflux, an

indicator of early apoptosis, induced by the release of potent pro-apoptotic signalling

molecules such as Cytochrome C and Apoptosis Inducing Factor (AIF) (Juhaszova et al. 2004).

Figure 26 demonstrates the mitochondrial membrane potential change initiated after 24 and

48 hours of 10µM and 30µM phenoxodiol treatment. FCCP positive controls gave an indication

of maximal potential depolarisation rate and were significantly different from all control,

10µM and 30µM phenoxodiol treatments across all cell lines (p<0.001). LNCaP cells displayed a

significant decrease in membrane potential, versus control population, after 24 hours

treatment with 30µM treatment (p=0.012) and with both 10µM and 30µM treatments, versus

control, after 48 hours (p= 0.017, p<0.001). In the 48 hour treatment cohort the 30µM

phenoxodiol group were observed to have almost depolarised as much as the FCCP positive

control, establishing that LNCaP cells had a large increase in mitochondrial depolarisation,

indicative of a susceptibility to apoptotic signalling induced by phenoxodiol treatment. As a cell

line representative of early stage prostate cancer, this was an expected outcome after the lack

of response observed using the MTS absorbance assay.

DU145 mitochondrial membrane potential was significantly decreased at the 24 hour time

point in response to 30µM phenoxodiol treatment versus control (p<0.001) and 10µM

(p=0.0148). Over a 48 hour treatment course the 10µM concentration was significantly

depolarised versus control (p=0.0015) while the 30µM concentration was significantly different

to control (p<0.001) and 10µM (p=0.00485). The DU145 cell response to treatment indicated a

susceptibility to apoptotic signalling in a dose and time dependent manner with 48 hour

treatment time being more effective than the 24 hour time point. DU145 cell depolarisation,

after phenoxodiol treatment, never approached the maximal depolarisation rate indicated by

the FCCP treatments. The comparison indicated a reduced sensitivity to mitochondrial

depolarisation in response to phenoxodiol treatment and therefore reduced early apoptotic

Page 144: Investigating the effects of, the isoflavone, Phenoxodiol

124

signalling. Resistance to apoptotic induction is typical from a cell line indicative of late stage

prostate cancer.

Finally, PC3 mitochondrial membrane potential was significantly decreased at the 24 hour time

point in response to 10µM Phenoxodiol treatment versus control (p=0.03) while the 30µM

treatment was significantly different from control (p=0.0121) and 10µM PXD (p=0.0215). Post

48 hours of treatment the 10µM concentration was significantly different to control (p=0.018)

while the 30µM concentration significantly different to control (p=0.0024) but not to 10µM

(p=0.06). The PC3 cell mitochondrial depolarisation rate in response to treatment was even

less than that of DU145 cells which was expected from the cell line that represents a more

aggressive late stage prostate cancer etiology. Though there was significant increase in PC3

mitochondrial depolarisation versus control, the small overall increase indicated that PC3 cells

were not as susceptible to early apoptotic signalling as either LNCaP or DU145 cells.

Page 145: Investigating the effects of, the isoflavone, Phenoxodiol

125

FIGURE 26 JC-1 ANALYSIS OF MITOCHONDRIAL MEMBRANE POTENTIAL OVER 24 AND 48 HOURS OF

TREATMENT

* indicates significance relative to control for that time point. ** indicates significance relative

to 10µM phenoxodiol treatment and control for that time point. *** indicates significance

0

20

40

60

80

100

120

24 Hours 48 HoursFlo

ure

sce

nce

Re

lati

ve t

o C

on

tro

l

Treatment

LNCaP JC-1 Analysis

Control

10µM PXD

30µM PXD

FCCP

0

20

40

60

80

100

120

24 Hours 48 HoursFlo

ure

sce

nce

Re

lati

ve t

o C

on

tro

l

Treatment

DU145 JC-1 Analysis

Control

10µM PXD

30µM PXD

FCCP

0

20

40

60

80

100

120

24 Hours 48 HoursFlo

ure

sce

nce

Re

lati

ve t

o C

on

tro

l

Treatment

PC3 JC-1 Analysis

Control

10µM PXD

30µM PXD

FCCP

Page 146: Investigating the effects of, the isoflavone, Phenoxodiol

126

relative to 30µM phenoxodiol and control for that time point. Cell depolarisation is the

green/red fluorescence ratio relative to the control.

Caspase-3 is a potent stimulator of apoptosis and the final effector caspase of the caspase

cascade apoptosis pathway. Caspase-3 exists as the constant target of extrinsic and intrinsic

apoptotic cell signalling and is a direct target of anti-apoptotic cell signalling (Janicke et al.

1998; Boatright and Salvesen 2003)). Figure 27 demonstrates activated Caspase-3 expression

measured through fluorescence analysis after 24 and 48 hours of phenoxodiol treatment.

LNCaP Caspase-3 activity was found to significantly decrease after 48 hours treatment with

10µM phenoxodiol treated cells experiencing a significant decrease in Caspase-3 expression

versus DMSO vehicle control (p=0.003) and 30µM PXD treated cells experiencing significant

expression decrease versus control (p<001) and 10µM (p=0.043). The total fluorescence

decrease measured was 18.1% (10µM) and 29.08% (30µM) respectively. DU145 cells had a

small but significant increase in Caspase-3 expression over 48 hours with both 10µM and

30µM (p=0.009, p=0.0123) phenoxodiol treatments expressing more activated Caspase-3. The

measured fluorescence increase was only 4.5% (10µM) and 3.8% (30µM) over the control

population. Such a small increase could potentially have biologically significant effects

however in view of the lack of response from PC3 cells and the decreased expression by LNCaP

cells it might be viewed that this small change, while statistically significant, is not the driving

factor in cell cytotoxicity in DU145 cells. PC3 cells had no significant changes in activated

Caspase-3 expression measured over 24 and 48 hours after 10µM and 30µM phenoxodiol

treatment even though there was a clear anti-proliferative effect demonstrated in the MTS

assay. This data suggests, but does not confirm, that phenoxodiol induces cell cytotoxicity

through a caspase independent manner.

Page 147: Investigating the effects of, the isoflavone, Phenoxodiol

127

FIGURE 27 FLUORESCENT ANALYSIS OF CASPASE-3 ACTIVITY AFTER PHENOXODIOL TREATMENT OVER

24 AND 48 HOURS

* indicates significance relative to control for that time point. ** indicates significance relative

to 10µM Phenoxodiol treatment and control for that time point. A set gain level between cell

lines and time points was used.

05000

1000015000200002500030000350004000045000

24 Hours 48 Hours

Ab

sorb

ance

52

0n

m

Time

LNCaP Caspase-3 Activity

Control

10µM PXD

30µM PXD

05000

1000015000200002500030000350004000045000

24 Hours 48 Hours

Ab

sorb

ance

52

0n

m

Time

DU145 Caspase-3 Activity

Control

10µM PXD

30µM PXD

05000

1000015000200002500030000350004000045000

24 Hours 48 Hours

Ab

sorb

ance

52

0n

m

Time

PC3 Caspase-3 Activity

Control

10µM PXD

30µM PXD

Page 148: Investigating the effects of, the isoflavone, Phenoxodiol

128

Apoptotic DNA fragmentation is a key feature of apoptosis, and is achieved by caspase

activation of endogenous endonucleases which subsequently cleave chromatin DNA into

fragments of around 180 base pairs multiples thereof. The amount of laddering can be used as

a direct measure of the amount of induced apoptosis (McCarthy et al. 1997). Figure 28

demonstrates DNA fragmentation through the quantitative measurement of three prime end

labelling (3’-End Labelling) which measures fragmentation levels by scintillation count.

Laddering can also indicate an apoptotic versus necrotic response and Figure 29 demonstrates

the qualitative measure of 3’-End labelling viewed through exposure to film. In the LNCaP cells

DNA laddering was clearly evident while in the DU145 samples laddering and a slight smear

was visible. PC3 cells had a necrotic smear in the film and no evidence of laddering. Due to the

measured response of PC3 as being necrotic, only the LNCaP and DU145 samples were

measured in the liquid scintillation counter to qualitatively assess their total apoptotic

response. LNCaP cells displayed a significant increase in apoptotic laddering versus control

after 24 hours of treatment with both 10µM (p=0.013) and 30µM (p=003) treatments,

exhibiting increased apoptosis by 35% (10µM) and 39% (30µM) respectively, compared to

control. After 48 hours of treatment with 30µM phenoxodiol the LNCaP cells were determined

to have a statistically significant increase in apoptosis versus control (p=0.044) exhibiting a 63%

increase in apoptotic laddering.

This result confirms the LNCaP cells were sensitive to apoptosis induction as indicated by JC-1

analysis. DU145 cells had a significant increase in apoptosis over 24 hours in response to the

30µM phenoxodiol treatment versus control (p=0.02) and versus 10µM (p=0.017) with an

increase in percentage of apoptosis by 15% versus control. After 48 hours of treatment it was

observed that 10µM phenoxodiol treatment exhibited increased apoptosis versus control

(p=0.009) and a 66% increase in apoptosis. The 30µM phenoxodiol treatment significantly

increased apoptosis versus control (p=0.03) and 10µM phenoxodiol (p<0.05), after 48 hours,

Page 149: Investigating the effects of, the isoflavone, Phenoxodiol

129

with a 700% increase in the 30µM treatment dose versus control indicating that DU145 cells

were very sensitive to apoptosis induced by phenoxodiol treatment after 48 hours of

treatment.

FIGURE 28 3'-END LABELLING APOPTOTIC ANALYSIS POST PHENOXODIOL TREATMENT OVER 24 AND 48

HOURS

* indicates significance relative to control for that time point. ** indicates significance relative

to 10µM Phenoxodiol treatment and control for that time point. Counts Per Minute (CPM) is

the amount of low molecular weight DNA labelled with radioactive P32 by the ddATP enzyme.

([α32P]-ddATP).

0

500

1000

1500

2000

2500

24 Hours 48 Hours

Low

MW

DN

A L

abe

llin

g (C

PM

)

Treatment

LNCaP 3'-End Labelling

Control

10µM PXD

30µM PXD

0

1000

2000

3000

4000

5000

6000

24 Hours 48 Hours

Low

MW

DN

A L

abe

llin

g (C

PM

)

Treatment

DU145 3'-End Labelling

Control

10µM PXD

30µM PXD

Page 150: Investigating the effects of, the isoflavone, Phenoxodiol

130

FIGURE 29 AN EXAMPLE OF LNCAP 3’-END LABELLING QUALITATIVE DNA LADDERING AFTER EXPOSURE

TO FILM

The final functional measurement of apoptosis was executed using an Annexin-V-

Fluos/Propidium Iodide (AV/PI) double stain on live cells which were placed into a FACSCalibur

flow cytometer for the measurement of cell fluorescence. No staining indicated normal cell

function, Annexin-V staining indicated early stage apoptosis by the conjugation of Annexin-V to

membrane bound, but externalised phosphatidylserine. Annexin-V only staining represents

early apoptosis as the cells still retain the ability to exclude PI. The AV/PI double stain indicated

late stage apoptosis/necrosis as the AV was bound to externalised phosphatidylserine and the

PI was bound to nuclear DNA which the cells lacked the ability to exclude (Martin et al. 1995).

Figure 30 demonstrates the cytotoxic effects of 24 and 48 hours of 10µM and 30µM

phenoxodiol treatment, as measured by fluorescence activated cell sorting (FACS) analysis also

called flow cytometry.

After 24 hours of treatment LNCaP cells indicated no significant changes in relation to control

populations; however, a high level of early and late apoptotic cell populations was present.

DMSO 10µM 30µM

Page 151: Investigating the effects of, the isoflavone, Phenoxodiol

131

After 48 hours of treatment the 10µM phenoxodiol treated cells were significantly different

versus control in the alive (p<0.0008, 64% decreased), early apoptosis (p=0.014, 153%

increase) and late apoptosis (p=0.0005, 332%) cell populations. The 48 hour 30µM

phenoxodiol treatments were significantly different from control in the alive (p<0.05, 50%

decrease) and late apoptotic population (p=0.03, 216% increase) but the early apoptotic

population was not statistically significant from control early apoptotic (p=0.130). LNCaP cells

were initially unresponsive then became very sensitive to phenoxodiol treatment, with the live

cell population decreasing and the early and late apoptotic populations increasing significantly

in response to treatment. The LNCaP cell population were difficult to obtain a 10,000 viable

cell count for each sample due to phenoxodiol’s cytotoxicity.

DU145 cells had significant changes over 24 hours in both 10µM and 30µM phenoxodiol

treatments. The 10µM treatment significantly decreased alive cells versus control (p=0.037,

12.3% decrease), significantly increased early apoptotic (p=0.047, 74% increase) and

significantly increased late apoptotic (p=0.048, 100% increase) cell populations. The 30µM

treatment group had a significant early apoptotic population increase versus control

(p=0.0006, 291% increase) and difference versus 10µM phenoxodiol treatment in early

apoptotic (p=0.003, 125% increase) and late apoptotic (p=0.02, 58% decrease) cell populations.

Over 48 hours the 10µM treatment group exhibited significantly decreased live cell population

(p=0.0002, 17% decrease), significantly increased early apoptotic (p=0.0002, 3000% increase)

and late apoptotic (p=0.0026, 69% increase) cell populations versus control. The 30µM

treatment group was significantly different from both control and 10µM treatments at live cell

population (p<0.001, 37% decrease, p<0.0004, 24.1% decrease), early apoptotic (p<0.001,

5000% increase, p=0.0007, 75% increase) and late apoptotic (p=0.0002, 251% increase,

p=0.0008, 108% increase) cell populations. The DU145 cells increased both early and late

apoptotic populations in a time/dose dependant manner and reflected the decreasing cell

Page 152: Investigating the effects of, the isoflavone, Phenoxodiol

132

viability. Phenoxodiol treatment of DU145 cells did not exhibit as large a decrease in live cell

population after treatment as was observed in LNCaP cells, but was still able to induce

significant cytotoxicity in the prostate cancer cell line.

PC3 cells indicated no significant changes in early apoptotic cell population across the 24 and

48 hour treatment time points. After 24 hours of treatment with 10µM phenoxodiol treatment

a significant difference versus control in both live cell (p=0.0007, 15% decrease) and necrotic

(p=0.0001, 180% increase) was exhibited while the 30µM treatment had a significant

difference versus control in both live cell (p=0.0001, 23.2% decrease) and necrotic (p=0.0007,

272% increase) and a significant difference versus 10µM in necrotic (p=0.03, 33% increase) cell

population. The 48 hour 10µM treatment group had a significant difference versus control in

both live cell (p<0.001, 24.6% decrease) and necrotic (p<0.001, 287% increase) while the 30µM

treatment had a significant difference versus control in both live cell (p<0.001, 34.1%

decrease) and necrotic (p<0.001, 415% increase) and a significant difference versus 10µM in

live cell (p=0.0004, 12.7% decrease) and necrotic (p<0.001, 32.9% increase) cell population.

During 3’-end labelling it was determined that PC3 cells responded to phenoxodiol by a

necrotic response. The results of the AV/PI FACS analysis correlate with only the late stage

apoptotic/necrotic double staining cells increasing in number, indicative of a purely necrotic

response. PC3 cells were less susceptible to phenoxodiol cytotoxicity versus the other cell

lines, with a 34.1% decrease in live cell population after 48 hours of treatment with a 30µM

dose, this is compared to the 37% decrease in viable cell numbers from DU145 and 50%

decrease in LNCaP cells.

Page 153: Investigating the effects of, the isoflavone, Phenoxodiol

133

FIGURE 30 ANNEXIN V-FLUOS / PROPIDIUM IODIDE DOUBLE STAINING ANALYSIS OF PROSTATE CANCER

CELLS POST PHENOXODIOL TREATMENT OVER 24 AND 48 HOURS

0

20

40

60

80

100

Control 10µM PXD24 Hours

30µM PXD Control 10µM PXD48 Hours

30µM PXD

% o

f C

ells

Treatment

LNCaP AV/PI FACS Analysis

Alive Early Apoptotic Late Apoptotic / Necrotic

0

20

40

60

80

100

Control 10µM PXD24 Hours

30µM PXD Control 10µM PXD48 Hours

30µM PXD

% o

f C

ells

Treatment

DU145 AV/PI FACS Analysis

Alive Early Apoptotic Late Apoptotic/Necrotic

0

20

40

60

80

100

Control 10µM PXD24 Hours

30µM PXD Control 10µM PXD48 Hours

30µM PXD

% o

f C

ells

Treatment

PC3 AV/PI FACS Analysis

Alive Early Apoptotic Late Apoptotic/Necrotic

Page 154: Investigating the effects of, the isoflavone, Phenoxodiol

134

* indicates significance relative to control group for that time point i.e. 24 hour Control early

apoptotic versus 24 hour 10µM Phenoxodiol early apoptotic. ** indicates significance relative

to 10µM Phenoxodiol treatment and control for that time point i.e. 48 hour Control 10µM

Phenoxodiol versus 48 Hour 30µM Phenoxodiol Control.

4.5. DISCUSSION

Determining appropriate initial cell concentration rates ensured logarithmic growth after 96

hours and negated any cellular activity decrease due to cell crowding effects. We demonstrate

here, utilizing the MTS viability assay, that phenoxodiol elicits time- and dose-dependent anti-

proliferative activity against both androgen-responsive and androgen resistant prostate cancer

cell lines. Initial results from the proliferation assay indicated LNCaP was less susceptible to

phenoxodiol treatment than either DU145 or PC3 cells; however, visualising the cells indicated

the opposite, a majority of LNCaP cells, far more than DU145 or PC3, had been induced into

cytotoxicity and were dead. As we had already accounted for the interaction between

Phenoxodiol and MTS dye there was no indication as to why the MTS assay suggested very

little decrease in proliferative activity yet visualisation of the cells clearly indicated significant

cytotoxicity was occurring. (Wang et al. 2010) has reported an underestimating of anti-

proliferative effects of cytotoxic compounds due to MTS dye in LNCaP cells than when

compared to other techniques. While the LNCaP proliferation assay was not necessarily

sensitive to the MTS addition, the use of apoptotic assays, 3’end labelling, JC-1 and AV/PI flow

cytometry, ensured accurate cytotoxicity information was obtained. DU145 and PC3 cell lines

indicated dose and time responsive cytotoxicity over 24 and 48 hours treatment with

phenoxodiol. While both cell lines are considered examples of late stage prostate cancer, PC3

is acknowledged as an extremely chemoresistant cell line and correspondingly had the least

decrease in proliferation.

Page 155: Investigating the effects of, the isoflavone, Phenoxodiol

135

Acidity measurements in cell culture media resulted in no significant pH changes detected

between control and treatment groups using total media at 37°C. Variation in cell cytotoxicity

and proliferation decreases were not due to pH changes in the cell media. JC-1 mitochondrial

depolarisation assays indicated depolarisation under the influence of phenoxodiol treatment.

Mitochondrial catastrophe is a potent inducer of cell cytotoxicity and considered an early

apoptotic indicator as it stimulates the intrinsic caspase pathway via the conformational

change of the pro-apoptotic Bcl-2 family members. This causes mitochondrial efflux of more

potent pro-apoptotic Bcl-2 family members such as Bax, resulting in an imbalance between

anti- and pro-apoptotic signalling (Gross et al. 1999; Kuruvilla et al. 2003). FCCP was used as

the positive control and indicated the maximum potential depolarisation of the cells at each

time point and treatment (Kuruvilla et al. 2003). LNCaP cells were very susceptible to cell

death signalling, with 30µM phenoxodiol treatment almost equalling maximal mitochondrial

depolarisation indicated by FCCP. As the cell line indicative of an early stage prostate cancer,

the LNCaP response to phenoxodiol as detected using the JC-1 assay was more accurate than

the MTS assay. DU145 and PC3 both had significant depolarisation, although far more limited

than LNCaP compared to FCCP depolarisation. This was another indication the models of late

stage prostate cancer, known for their chemoresistance, were appropriate choices for this

study. There is currently no curative treatment for late stage prostate cancer (Canfield et al.

2006).

Classic apoptosis, such as that induced by chemotherapeutic agents, can proceed via extrinsic

death-receptor mediated and intrinsic mitochondrial-mediated pathways ultimately resulting

in the activation of Caspase-3 (Asselin et al. 2001; Mahoney 2007). Given that phenoxodiol

induced classic apoptosis in ovarian cancer (Gamble et al. 2005; Alvero et al. 2006), it was

surprising to find that activated Caspase-3 expression remained unchanged in PC3 cells and

even declined in LNCaP cells, while the actual increase in expression in DU145 cells was

Page 156: Investigating the effects of, the isoflavone, Phenoxodiol

136

minimal. Activated Caspase-3 activity assays indicated that phenoxodiol may function through

a caspase independent pathway, as LNCaP levels of activated Caspase-3 were significantly

decreased after 48 hours of phenoxodiol treatment while the cells were undergoing detectable

apoptosis. The lack of consistent signalling indicated that cell cytotoxicity induced by

phenoxodiol was potentially initiated via a caspase independent cell death pathway.

Apoptotic analysis by 3′-end labelling and Annexin-V/PI based flow cytometry validated both

qualitatively and quantitatively that phenoxodiol induced apoptotic cell death in LNCaP and

DU145 cell lines. Apoptosis was detected by DNA fragmentation analysis and Annexin-V

binding to the externalisation of phosphatidylserine only, indicating early stage apoptosis; or in

conjunction with propidium iodide nuclear staining, indicating late stage apoptosis. In contrast,

phenoxodiol significantly decreased the proliferation of PC3 cells in the absence of DNA

laddering and externalisation of phosphatidylserine, indicating a necrotic response that was

qualitatively measured by 3’-end labelling and thereby confirming the response. The use of

only one apoptotic measure, the lack of confirmation of DNA laddering or the use of only

propidium iodide in flow cytometry makes singular technique apoptotic studies flawed.

Annexin-V phosphatidylserine only and AV/PI double staining is required to detect early versus

late apoptosis or necrosis but does not guarantee either. Confirmation with DNA laddering

ensures the outcome with a quick technique that is both qualitative and quantitative, but by

itself DNA laddering cannot indicate early stage apoptosis. FACS analysis established that

phenoxodiol clearly induced; the appearance of early and late stage apoptotic populations in

LNCaP and DU145 cells, the appearance of necrosis in PC3 cells as well as a marked decrease in

viable cells in all cell lines. These data were qualitatively confirmed by the visualisation of

characteristic low molecular weight laddering during 3’-end labelling. All three cell lines

responded to phenoxodiol treatment with a minimum decrease in viable cells of 30% after

only 48 hours treatment, and a variable incidence of early apoptotic and late stage

Page 157: Investigating the effects of, the isoflavone, Phenoxodiol

137

apoptotic/necrotic cell population production. In all modes of analysis, cell death was reliant

on dose and duration of phenoxodiol exposure and yet Caspase-3 was not activated in

phenoxodiol treated PC3 cells. Neither LNCaP nor DU145 cells had any consistent caspase 3

change observed which implicated a non caspase dependant route as the target of apoptotic

induction.

Phenoxodiol exhibits significant cytotoxicity, inducing cell death in the prostate cancer cell

lines LNCaP, DU145 and PC3. All three cell lines exhibited significantly decreased viable cell

populations after only 48 hours of treatment but they also utilised different signalling

pathways than those reported in previous studies (Kamsteeg et al. 2003; Sapi et al. 2004;

Alvero et al. 2006; Mor et al. 2006). This establishes the cell-death inducing capabilities of

phenoxodiol on prostate cancer cells and suggests the induction of apoptotic cell death

through non-classical pathways. By targeting non classical apoptotic pathways and successfully

inducing cell cytotoxicity even in the most chemoresistant cell lines, phenoxodiol displays

potential as a drug for future treatment of prostate cancer.

Page 158: Investigating the effects of, the isoflavone, Phenoxodiol

138

5 . C H A P TER F IV E : C EL L D EA TH S IGNA L L IN G IN RES P ONS E T O

P H ENOX O D IOL TREA TMEN T

5.1. INTRODUCTION

Two of the hallmarks of cancer initiation and progression are the ability of cells to become

insensitive to anti-growth signals and to evade apoptosis. To do this the cells accumulate

increased expression of anti-apoptotic signalling molecules, which prevent activation of the

intrinsic and extrinsic signalling pathways that normally induce cell death (Hanahan and

Weinberg 2000). The accumulation of these factors, coupled with a dysregulation in the cell

cycle, results in a progressively increasing rate of mutation and the movement of cells from

benign to metastatic types. In response to cytotoxic insult from both external and internal

sources, cells will normally initiate machinery, such as caspases, that execute apoptotic cell

death.

Apoptosis is important as it results in cell death with DNA cleavage into unusable fragments

with no resultant inflammatory response. Achieving apoptosis results in fewer side effects

from chemotherapeutic agents and, as such, initiating an apoptotic response becomes the end

point target of all new therapy trials (Brown and Wouters 1999). Though they vary between

cells, the array of apoptosis induction signals commonly trigger signalling pathways that

coalesce at the mitochondria. Once stimulated, the mitochondrial cell walls efflux powerful

apoptotic stimulators such as Cytochrome c and Bax (Lee et al. 2008) to activate central

effector pathways involving a series of proenzymes, the caspases (Maguire et al. 2000). When

activated, caspases are efficient at degrading cellular processes and DNA to the point where

normal physiological activity is impossible, resulting in cellular response and apoptotic

Page 159: Investigating the effects of, the isoflavone, Phenoxodiol

139

phenomena (Hengartner 2000). The mitochondrial pathway of cell death is mediated by the

Bcl-2 family of proteins, a group of anti-apoptotic and pro-apoptotic proteins that regulates

the passage of small molecules such as Cytochrome c, Apoptosis Inducing Factor (AIF) and Bax,

which activates caspase cascades through the mitochondrial transition pore (Korsmeyer 1992;

Korsmeyer et al. 2000). The activation of caspases is counteracted by anti-apoptotic molecules

of the Bcl-2 family (Bcl-2, Bcl-xL), because these Bcl-2 family proteins heterodimerise with pro-

apoptotic members of the Bcl-2 family (Bax, Bak) and interfere with release of Cytochrome c

by inactivating pore-forming proteins (Gross et al. 1999). Cancer cells commonly have the anti-

apoptotic members of the Bcl-2 family increased in expression, allowing them to prevent

extrinsic anti-growth signalling and avoid intrinsic stimulation of apoptosis (Cory and Adams

2002).

Upon intrinsic stimulation of apoptosis, the release of Bax and Cytochrome c from the

mitochondria initiates the apoptosis promoting activity of Cytochrome c by allowing

interaction with the Apaf-1 protein. Binding of Cytochrome c to Apaf-1 causes recruitment of

Caspase-9, forming the apoptosome (Slee et al. 1999; Srinivasula et al. 2001; Zou et al. 2003).

The apoptosome is a large molecule that interacts with the executioner caspases, Caspase-3

and Caspase-7, which initiate DNA fragmentation and cause the cell to enter into apoptosis

(Boatright and Salvesen 2003). Caspase-3 can be activated via extrinsic pathways that produce

activated Caspase-9 through Caspase-8 signalling. If intrinsic pathways are activated, the

release of Cytochrome c and apoptosome formation results in Caspase-3 activation (Riedl and

Salvesen 2007). During normal cell function Caspase-3 is in an inactive zymogen form. The

zymogen feature of Caspase-3 is necessary because, if unregulated, caspase activity would kill

cells indiscriminately and therefore Caspase-3 activity is strictly limited until it is cleaved by an

initiator caspase after apoptotic signalling (Donepudi and Grutter 2002; Boatright and Salvesen

2003).

Page 160: Investigating the effects of, the isoflavone, Phenoxodiol

140

The Bcl-2 family members are not the only regulatory proteins of apoptosis nor are the

caspases the only method of executing programmed cell death. Members of the Inhibitors of

Apoptosis Proteins (IAP) such as xIAP prevent apoptosis by inactivating Caspase-3 and Caspase-

9 directly, amongst other roles, and are found in increased expression in chemoresistant

cancer cells (Dohi et al. 2004; Dubrez-Daloz et al. 2008). FLICE inhibitory protein (FLIP) binding

to Caspase-8 can also prevent death receptor mediated apoptosis and has been linked to

phenoxodiol treatment (De Luca et al. 2008). Apoptosis inducing factor (AIF) is involved in

initiating a caspase-independent pathway of apoptosis and can act as an NADH oxidase.

Normally it is found behind the outer membrane of the mitochondria and is therefore

secluded from the nucleus; however, when the mitochondria are damaged, it moves to the

cytosol and to the nucleus. Another AIF function is to regulate the permeability of the

mitochondrial membrane upon apoptosis stimulation, causing caspase dependant pathways to

be activated (Xie et al. 2005). Apoptosis can still happen in cells that lack Cytochrome c, Apaf-1

or caspases with AIF causing apoptotic changes independent of caspase activity (Susin et al.

1999; Li et al. 2000).

X-linked inhibitor of apoptosis protein (xIAP) is part of a family of endogenous caspase

inhibitors that, when up regulated, have been implicated in the early stage development of

prostate and breast epithelial carcinoma (Parton et al. 2002; Krajewska et al. 2003). In

conjunction with anti-apoptotic Bcl-2 members (Bcl-2, Bcl-xL) IAP family members exert

significant anti-apoptotic effects through; inhibition of Caspase-3 and -7 activation and

function, inactivation of pro-apoptotic Bcl-2 family members and involvement in signalling

pathways NF-κB and JNK (Levkau et al. 2001; Sanna et al. 2002; Hu et al. 2003; Dubrez-Daloz et

al. 2008). XIAP expression levels are also found to be increased in epithelial ovarian and

prostate cancers, and have been linked to chemoresistance and progression of cells to

metastasis (Berezovskaya et al. 2005).

Page 161: Investigating the effects of, the isoflavone, Phenoxodiol

141

Separate to direct apoptotic interaction, cytotoxic compounds can cause reactive oxygen

species production, which leads to oxidative stress in cancer cells by rapid accumulation of

highly reactive molecules such as nitric oxide (Murphy 2009). When this occurs it results in

damage to cell structures and potentially activates the apoptotic cascade making it a target for

chemotherapy agents (Dharmarajan et al. 1999). Oxidative stress induces the expression of

antioxidant genes that contain an antioxidant response element in their promoters; this can

result in a triggering of apoptosis or necrosis of various cell types and, in several instances, can

inhibit cell growth and interfere with the cell cycle (Minelli et al. 2009).

After determining that phenoxodiol induced cytotoxicity in prostate cancer cells, the next

focus was to determine the molecular signalling that resulted in cytotoxicity after phenoxodiol

treatment. Molecular signalling that results in apoptosis is generally initiated via extrinsic and

intrinsic pathways that mediate mitochondrial permeability and Bax upregulation. This results

in Caspase-3 activation and subsequent apoptosis. To elucidate the specific mechanism by

which phenoxodiol elicits an anti-proliferative effect in prostate cancer cell lines, we

investigated the interactions of the apoptotic signalling molecules AIF, Bax, Bcl-xL, Caspase-3

and xIAP in prostate cancer cells treated with phenoxodiol. Subsequently this chapter

investigates if caspase signalling is a requirement for phenoxodiol activity and whether

phenoxodiol treatment could induce the production of reactive oxygen species which would

then induce a cytotoxic response via oxidative stress.

Page 162: Investigating the effects of, the isoflavone, Phenoxodiol

142

5.2. AIMS

The aims of this chapter were as follows:

Aim 1: To determine consistency in cell death induction signalling between the cell lines by

investigation of AIF, Bax, Bcl-xL, Caspase-3 and xIAP expression changes post phenoxodiol

treatment.

Aim 2: To determine if phenoxodiol activity required intrinsic or extrinsic caspase signalling

pathways in LNCaP, DU145 and PC3 cells.

Aim 3: To determine if Phenoxodiol-induced reactive oxygen species generation by

investigating nitric oxide production post phenoxodiol treatment.

5.3. METHODOLOGY

The methodology utilised in this chapter is discussed in detail in the materials and methods

chapter (page 50). Quantitative PCR (qPCR) analysis and Western blot analysis were performed

on all three cell lines LNCaP, DU145 and PC3, after 24 and 48 hours exposure to 10µM and

30µM phenoxodiol concentrations. Briefly, cells were seeded into 6 well plates and treated

appropriately with phenoxodiol, RNA extraction was performed using the Chomczynski

guanidinium thiocyanate-phenol-chloroform method via a TRI Reagent® extraction kit

(Chomczynski and Sacchi 1987). Purified RNA was analysed for concentration using a

Nandrop™ light spectrophotometer and 1µg was firstly DNAse treated and then converted into

cDNA using a “M-MLV Reverse Transcriptase RNase H Minus, Point Mutant Taq” Promega

enzyme kit with random primers. Removal of excess random primers, reagent and salts was

performed using a Mol-Bio spin column post-pcr purification kit, leaving a final volume of 50µL

cDNA. Each qPCR experiment was performed using 2µL of cDNA per sample. Primers were

Page 163: Investigating the effects of, the isoflavone, Phenoxodiol

143

designed to test for AIF, Bax, Bcl-xL, Caspase-3 and xIAP by using published primers. Primer

spanning of an intron/exon boundary was confirmed using Primer3 and BLAST search engines.

Quantitative PCR (qPCR) was performed by adding 2µL of cDNA to a master mix of primers,

Taq polymerase, dNTP’s, salt and water to bring the final volume to 10µL. qPCR was then

performed on samples for 35 temperature cycles at varying annealing temperatures, on a

Corbett Rotorgene 3000 or 6000, with initial results run against a gel ladder to determine if

primer dimerisation had occurred. Once a successful sample was gel loaded and confirmed, a

gel extraction was performed, purifying the final product that was then serially diluted 10 fold

each step to produce a set of standards. In the full sample qPCR experiments, each sample was

duplicated and compared to five standards which were also duplicated. The standards were 10

fold dilutions of given concentrations and the resulting standard curve then determined final

sample concentrations. Final sample concentrations were compared against the expression of

L19.

Protein level analysis was performed using a Western blot protocol. Briefly, cells were seeded

into six well plates, treated with phenoxodiol appropriately and a RIPA buffer/β-

mercaptoethanol-based, whole cell lysate, protein extraction performed. Protein

concentration was determined against a standard curve on a Nanodrop™ light

spectrophotometer and then, with varying concentrations per protein to be detected, loaded

onto a SDS-Page/acrylamide based gel and current applied to the gel. After protein separation

through the SDS-Page/acrylamide gel the proteins were transferred to a nitrocellulose

membrane by the application of voltage to a transfer buffer. Ponceau S was used to stain the

nitrocellulose membrane to determine if transfer was effective, if air bubbles existed or if even

protein loading had occurred. Finally, primary antibodies were added to the membrane and

Page 164: Investigating the effects of, the isoflavone, Phenoxodiol

144

incubated in the appropriate conditions, excess was removed with TBS-T washes and a

secondary antibody conjugated to HRP applied in the appropriate conditions. A

chemiluminescent substrate kit was used to detect antibody binding concentration.

A caspase inhibition assay was performed using the pan-caspase inhibitor Z-VAD-FMK. Briefly,

cells were seeded into 96 well plates and after 48 hours of growth treated with either; DMSO

vehicle control, 30µM phenoxodiol, 10µM Z-VAD-FMK pan caspase inhibitor or a combination

of both 30µM phenoxodiol and 10µM Z-VAD-FMK, in complete media over a period of 48

hours. After 48 hours of treatment 20µL of MTS dye was added to each well and the plates

incubated for 3 hours before the absorbance of light at 492nm was measured on a LabSystems

plate reader.

The Griess assay was performed to determine if phenoxodiol induced the production of nitric

oxide, a potent free radical. Cells were plated into 96 well plates and grown for 48 hours

before the media were replaced with a negative control, DMSO Control, 10µM phenoxodiol,

30µM phenoxodiol, 1000µM DEAN or 100µM SNP containing media. After 24 or 48 hours of

treatment a standard curve was performed on the plate with complete media and 10 fold

dilutions of a 100µM nitrite solution. After the standard curve was prepared 50µL of the

treatment media was aspirated from each well and replaced with 50µL sulfanilamide solution

and then incubated for 10 minutes. After 10 minutes there was an addition of 50µL of NED

solution to each well followed by another 10 minute incubation. After incubation the plate was

scanned on a LabSystems plate reader at 540nm absorbance to determine NO production rate.

Page 165: Investigating the effects of, the isoflavone, Phenoxodiol

145

5.4. RESULTS

Analysis of mRNA signalling expression by qPCR was performed to determine the extrinsic and

intrinsic apoptotic pathways activated by phenoxodiol treatment. The pro-apoptotic molecule

Apoptosis Inducing Factor (AIF) can induce apoptosis, in a caspase independent manner, by

causing chromatin condensation, interacting with NADH oxidase, inducing DNA fragmentation

and interacting with the mitochondrial membrane, thereby increasing permeability (Xie et al.

2005). Figure 31 demonstrates the quantitative mRNA expression of the AIF, over 24 and 48

hours, post phenoxodiol treatment. LNCaP AIF expression levels were significantly decreased

after 48 hours of treatment with 10µM (p=0.026) and 30µM (p=0.042) phenoxodiol

concentrations. At the 24 hour time point all cell lines expressed no significant difference in

comparison to the DMSO vehicle control levels. This result was also seen after 48 hours for the

DU145 and PC3 cell lines. AIF mRNA expression was determined to not be consistently altered

by phenoxodiol treatment between the prostate cancer cell lines.

Page 166: Investigating the effects of, the isoflavone, Phenoxodiol

146

FIGURE 31 AIF MRNA EXPRESSION ANALYSIS OF PROSTATE CANCER CELLS POST PHENOXODIOL

TREATMENT OVER 24 AND 48 HOURS

* indicates significance relative to control for that time point.

0

2

4

6

8

10

24 Hours 48 Hours

AIF

/L1

9

Treatment

LNCaP AIF qPCR Expression

Control

10µM PXD

30µM PXD

* *

0

1

2

3

4

5

6

7

8

24 Hours 48 Hours

AIF

/L1

9

Treatment

DU145 AIF qPCR Expression

Control

10µM PXD

30µM PXD

0

5

10

15

20

25

30

35

24 Hours 48 Hours

AIF

/L1

9

Treatment

PC3 AIF qPCR Expression

Control

10µM PXD

30µM PXD

Page 167: Investigating the effects of, the isoflavone, Phenoxodiol

147

Increased expression of the pro-apoptotic molecule Bax results in mitochondrial membrane

permeabilisation, releasing several pro-apoptotic molecules, such as Cytochrome c, which

induce apoptosis through activation of caspases (Cheng et al. 2001). Figure 32 demonstrates

the quantitative mRNA expression of pro-apoptotic molecule Bcl-2-associated X protein (Bax),

over 24 and 48 hours, post phenoxodiol treatment in prostate cancer cells. LNCaP and PC3

cells expressed no significant difference in Bax expression between DMSO vehicle control and

treatment groups over either time points. DU145 cells expressed an increase in Bax signalling

with both 10µM (p<0.05) and 30µM (p=0.037) phenoxodiol treatment groups after 48 hours

but no significant difference in the 24 hour cohort. There was no significant difference

detected between Bax signalling levels after treatment with either phenoxodiol

concentrations. Bax signalling was not found to be consistently altered across the prostate

cancer cells lines in response to phenoxodiol treatment.

Page 168: Investigating the effects of, the isoflavone, Phenoxodiol

148

FIGURE 32 BAX MRNA EXPRESSION ANALYSIS OF PROSTATE CANCER CELLS POST PHENOXODIOL

TREATMENT OVER 24 AND 48 HOURS

* indicates significance relative to control for that time point.

0

5

10

15

20

25

24 Hours 48 Hours

Bax

/L1

9

Treatment

LNCaP Bax qPCR Expression

Control

10µM PXD

30µM PXD

0

5

10

15

20

24 Hours 48 Hours

Bax

/L1

9

Treatment

DU145 Bax qPCR Expression

Control

10µM PXD

30µM PXD

*

0

5

10

15

20

25

24 Hours 48 Hours

Bax

/L1

9

Treatment

PC3 Bax qPCR Expression

Control

10µM PXD

30µM PXD

Page 169: Investigating the effects of, the isoflavone, Phenoxodiol

149

Bcl-xL protects cells from apoptosis by regulating mitochondria membrane potential and

volume by interacting with pro-apoptotic members, Bax or Bim, and subsequently preventing

the release of Cytochrome c and other mitochondrial factors from the intermembrane space

into the cytosol (Sun et al. 2008). Figure 33 demonstrates the quantitative mRNA expression of

the anti-apoptotic molecule Bcl-xL, over 24 and 48 hours, post phenoxodiol treatment in

prostate cancer cells. The expression levels of Bcl-xL were not found to change significantly in

response to treatment with Phenoxodiol at any time point. Bcl-xL mRNA signalling was

consistently unresponsive following exposure to phenoxodiol treatment.

Page 170: Investigating the effects of, the isoflavone, Phenoxodiol

150

FIGURE 33 BCL-XL MRNA EXPRESSION ANALYSIS OF PROSTATE CANCER CELLS POST PHENOXODIOL

TREATMENT OVER 24 AND 48 HOURS

0

5

10

15

20

25

24 Hours 48 Hours

Bcl

-xL/

L19

Treatment

LNCaP Bcl-xL qPCR Expression

Control

10µM PXD

30µM PXD

0

10

20

30

40

50

60

70

24 Hours 48 Hours

Bcl

-xL/

L19

Treatment

DU145 Bcl-xL qPCR Expression

Control

10µM PXD

30µM PXD

0

0.5

1

1.5

2

24 Hours 48 Hours

Bcl

-xL/

L19

Treatment

PC3 Bcl-xL qPCR Expression

Control

10µM PXD

30µM PXD

Page 171: Investigating the effects of, the isoflavone, Phenoxodiol

151

Caspase-3 is a potent downstream apoptotic signalling molecule that, once activated, can

induce DNA fragmentation and other apoptotic events (Janicke et al. 1998; Janicke et al. 1998).

Figure 34 demonstrates the quantitative mRNA expression of the pro-apoptotic molecule

Caspase-3 over 24 and 48 hours, post phenoxodiol treatment in prostate cancer cells. LNCaP

and PC3 cells were not found to significantly change expression of Caspase-3 signalling over

any course of treatment and time points. The 48 hours of treatment, 30µM PXD, PC3 sample

was not significant (p=0.25). After 24 hours treatment DU145 cells indicated no expression

differences over DMSO vehicle control but the 48 hour 10µM (p=0.023) and 30µM (p<0.05)

PXD treatments were found to be significantly increased over DMSO vehicle control. The

Caspase-3 mRNA signalling results corresponded to the Caspase-3 activity assay and indicated

increased expression of Caspase-3 in DU145 cells after 48 hours of treatment with

phenoxodiol. While activated Caspase-3 was seen to decrease (chapter 4) after 48 hours in

LNCaP cells, the signalling was not found to be altered. The previously reported decrease in

activated Caspase-3, but a lack of detected signalling change could be due to the dramatic

decrease in live cell population after 48 hours in the LNCaP cell line. Therefore the LNCaP cells

would need to be tested for activated Caspase-3 between 24 and 48 hours to determine the

period where Caspase-3 signalling and expression is detectably altered but a majority of the

LNCaP cell population has not progressed to late stage apoptosis.

Page 172: Investigating the effects of, the isoflavone, Phenoxodiol

152

FIGURE 34 CASPASE-3 MRNA EXPRESSION ANALYSIS OF PROSTATE CANCER CELLS POST PHENOXODIOL

TREATMENT OVER 24 AND 48 HOURS

* indicates significance relative to control for that time point.

0

10

20

30

40

50

24 Hours 48 Hours

Cas

pas

e-3

/L1

9

Treatment

LNCaP Caspase-3 qPCR Expression

Control

10µM PXD

30µM PXD

0

0.1

0.2

0.3

0.4

0.5

0.6

24 Hours 48 Hours

Cas

pas

e-3

/L1

9

Treatment

DU145 Caspase-3 qPCR Expression

Control

10µM PXD

30µM PXD

*

0

2

4

6

8

10

12

24 Hours 48 Hours

Cas

pas

e-3

/L1

9

Treatment

PC3 Caspase-3 qPCR Expression

Control

10µM PXD

30µM PXD

Page 173: Investigating the effects of, the isoflavone, Phenoxodiol

153

xIAP is a potent inhibitor of apoptosis that interacts with Caspase-3 thereby preventing

downstream apoptotic effects and Caspase-9 signalling (Straszewski-Chavez et al. 2004). Figure

35 demonstrates the quantitative mRNA expression of the anti-apoptotic molecule X-linked

Inhibitor of Apoptosis Protein (xIAP), over 24 and 48 hours, post phenoxodiol treatment in

cells. No significant changes in xIAP mRNA expression levels were detected across any

treatment group or time point. Phenoxodiol does not increase cytotoxicity through depression

of xIAP mRNA signalling in prostate cancer cell lines.

Page 174: Investigating the effects of, the isoflavone, Phenoxodiol

154

FIGURE 35 XIAP MRNA EXPRESSION ANALYSIS OF PROSTATE CANCER CELLS POST PHENOXODIOL

TREATMENT OVER 24 AND 48 HOURS

Western blot protein analysis was performed on whole cell lysate to determine protein level

changes after phenoxodiol treatment for the genes AIF, Bax, Bcl-xL and xIAP. Activated

0

0.5

1

1.5

2

2.5

3

3.5

24 Hours 48 Hours

xIA

P/L

19

Treatment

LNCaP xIAP qPCR Expression

Control

10µM PXD

30µM PXD

0

1

2

3

4

5

24 Hours 48 Hours

xIA

P/L

19

Treatment

DU145 xIAP qPCR Expression

Control

10µM PXD

30µM PXD

0

2

4

6

8

10

12

24 Hours 48 Hours

xIA

P/L

19

Treatment

PC3 xIAP qPCR Expression

Control

10µM PXD

30µM PXD

Page 175: Investigating the effects of, the isoflavone, Phenoxodiol

155

Caspase-3 protein was not detected because an activated Caspase-3 fluorescent activity assay

had already been performed. Cell protein expression was measured using densitometric

analysis after performing a Western blot, with protein expression compared to a standardising

protein, β-actin. AIF protein is a potent pro-apoptotic molecule that can induce apoptosis via a

caspase independent manner through mitochondrial permeabilisation (Susin et al. 1999).

Figure 36 demonstrates quantitative protein levels of the pro-apoptotic AIF protein over 24

and 48 hours post phenoxodiol treatment in prostate cancer cells. LNCaP cells were

determined to have significantly higher AIF protein levels after 24 hours of treatment with

30µM (p=0.015) PXD treatment, while at the 48 hour time point the 10µM PXD treatment was

found to express significantly lower levels of AIF than DMSO vehicle control (p=0.007) and

30µM PXD (p=0.008). The cell line DU145 was found to have no detectable levels of AIF

expression in any time point or group. PC3 cells were found to have significantly lower levels of

AIF protein versus DMSO vehicle control after 24 hours and 10µM of PXD treatment (p=0.036);

while, in the 48 hour treatment group, the 30µM PXD cohort exhibited a strong trend towards

biological significance but was not statistically significant (p=0.08). AIF protein expression was

not determined to consistently change after phenoxodiol treatment across the prostate cancer

cell lines.

Page 176: Investigating the effects of, the isoflavone, Phenoxodiol

156

FIGURE 36 AIF PROTEIN LEVEL ANALYSIS OF PROSTATE CANCER CELLS POST PHENOXODIOL TREATMENT

OVER 24 AND 48 HOURS

* indicates significance relative to control for that time point. ** indicates significance relative

to 30µM Phenoxodiol treatment and control for that time point.

Bax is an important pro-apoptotic protein essential for the mitochondrial release of

Cytochome c, inducing cellular apoptosis (Lee et al. 2008). Figure 37 demonstrates quantitative

protein levels of the pro-apoptotic Bax protein over 24 and 48 hours post phenoxodiol

treatment in prostate cancer cells. LNCaP cells were found to have significantly lower levels of

Bax protein after 24 hours of treatment with both 10µM (p=0.024) and 30µM (p≤0.001) PXD

treatment. No significant change was detected in the LNCaP 48 hour treatment group. DU145

cells did not express detectable levels of Bax protein, the second pro-apoptotic protein found

0.00

0.50

1.00

1.50

2.00

2.50

3.00

24 Hours 48 Hours

AIF

/β-a

ctin

Treatment

LNCaP AIF Protein Levels

Control

10µM PXD

30µM PXD

**

0

0.2

0.4

0.6

0.8

1

24 Hours 48 Hours

AIF

/β-a

ctin

Treatment

PC3 AIF Protein Levels

Control

10µM PXD

30µM PXD

*

Page 177: Investigating the effects of, the isoflavone, Phenoxodiol

157

to be down regulated to non-detectable levels in DU145 cells. PC3 cells expressed no

significant change in Bax protein expression across any time point or treatment group in

comparison to DMSO vehicle control. Bax protein expression was not determined to alter

consistently across the prostate cancer cell lines following treatment with phenoxodiol.

FIGURE 37 BAX PROTEIN LEVEL ANALYSIS OF PROSTATE CANCER CELLS POST PHENOXODIOL

TREATMENT OVER 24 AND 48 HOURS

* indicates significance relative to control for that time point.

Bcl-xL protein is a potent anti-apoptotic agent that acts through the binding of molecules with

BH3 domain-only molecules in stable mitochondrial complexes, thereby preventing activation

0

0.5

1

1.5

2

2.5

3

3.5

4

24 Hours 48 Hours

Bax

/β-a

ctin

Treatment

LNCaP Bax Protein Levels

Control

10µM PXD

30µM PXD*

*

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

24 Hours 48 Hours

Bax

/β-a

ctin

Treatment

PC3 Bax Protein Levels

Control

10µM PXD

30µM PXD

Page 178: Investigating the effects of, the isoflavone, Phenoxodiol

158

of pro-apoptotic pathways such as Bax (Cheng et al. 2001). Figure 38 demonstrates

quantitative protein levels of the anti-apoptotic Bcl-xL protein over 24 and 48 hours post

phenoxodiol treatment in prostate cancer cells. LNCaP cells expressed no significant

differences across 24 hours but, after 48 hours of treatment, both 10µM (p=0.003) and 30µM

(p=0.0099) PXD treatment groups were significantly increased in the expression of Bcl-xL

protein versus DMSO vehicle control. DU145 cells were found to have no significant expression

changes across 24 hours of treatment, although the 30µM PXD treatment group trended

towards biological significance (p=0.08) versus DMSO vehicle control. After 48 hours of

treatment the DU145 30µM PXD treatment group was found to be significantly different to

both the DMSO vehicle control (p=0.013) and the 10µM PXD treatment group (p=0.0018). PC3

cells expressed no significant differences between groups across 24 hours of treatment but,

after 48 hours of treatment, both 10µM (p=0.04) and 30µM (p=0.041) PXD treatment groups

had significantly lower expression of Bcl-xL protein than DMSO vehicle control. Bcl-xL protein

expression was altered in an inconsistent manner between the three cell lines following

phenoxodiol treatment.

Page 179: Investigating the effects of, the isoflavone, Phenoxodiol

159

FIGURE 38 BCL-XL PROTEIN LEVEL ANALYSIS OF PROSTATE CANCER CELLS POST PHENOXODIOL

TREATMENT OVER 24 AND 48 HOURS

* indicates significance relative to control for that time point. ** indicates significance relative

to 10µM Phenoxodiol treatment and control for that time point.

0

1

2

3

4

5

6

7

24 Hours 48 Hours

Bcl

-xL/

β-a

ctin

Treatment

LNCaP Bcl-xL Protein Levels

Control

10µM PXD

30µM PXD

*

*

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

24 Hours 48 Hours

Bcl

-xL/

β-a

ctin

Treatment

DU145 Bcl-xL Protein Levels

Control

10µM PXD

30µM PXD

**

0

0.1

0.2

0.3

0.4

0.5

24 Hours 48 Hours

Bcl

-xL/

β-a

ctin

Treatment

PC3 Bcl-xL Protein Levels

Control

10µM PXD

30µM PXD* *

Page 180: Investigating the effects of, the isoflavone, Phenoxodiol

160

X-linked inhibitor of apoptosis protein (xIAP) is part of a family of endogenous caspase

inhibitors that, when up regulated, have been implicated in the early stage development of

prostate and breast epithelial carcinoma (Parton et al. 2002; Krajewska et al. 2003). Figure 39

demonstrates quantitative protein levels of the anti-apoptotic xIAP protein over 24 and 48

hours post phenoxodiol treatment in prostate cancer cells. LNCaP cells were not found to have

a differential expression of xIAP protein across all time points and treatment groups versus

DMSO vehicle control. The DU145 cell population at 24 hours was determined to have a

significant decrease in xIAP protein expression in the 10µM (p=0.048) PXD versus DMSO

vehicle control, while the 30µM PXD treatment group was found to be significantly decreased

in comparison to both DMSO vehicle control (p=0.001) and 10µM PXD (p=0.005). No changes

were detected across the 48 hour treatment group in the DU145 cell population although the

30µM PXD group had a strong trend (p=0.07) towards a decrease in xIAP protein expression

compared to DMSO vehicle control. PC3 cells were not found to have a differential expression

of xIAP protein across all time points and treatment groups versus DMSO vehicle control.

Page 181: Investigating the effects of, the isoflavone, Phenoxodiol

161

FIGURE 39 XIAP PROTEIN LEVEL ANALYSIS OF PROSTATE CANCER CELLS OVER 24 AND 48 HOURS POST

PHENOXODIOL TREATMENT OVER 24 AND 48 HOURS

* indicates significance relative to control for that time point. ** indicates significance relative

to 10µM Phenoxodiol treatment and Control for that time point.

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

24 Hours 48 Hours

xIA

P/β

-act

in

Treatment

LNCaP xIAP Protein Levels

Control

10µM PXD

30µM PXD

0

0.2

0.4

0.6

0.8

1

24 Hours 48 Hours

xIA

P/β

-act

in

Treatment

DU145 xIAP Protein Levels

Control

10µM PXD

30µM PXD

*

**

0

0.5

1

1.5

2

24 Hours 48 Hours

xIA

P/β

-act

in

Treatment

PC3 xIAP Protein Levels

Control

10µM PXD

30µM PXD

Page 182: Investigating the effects of, the isoflavone, Phenoxodiol

162

Following the results of the Caspase-3 activity assay, qPCR and Western blots indicating that

the apoptotic signalling was inconsistent between cell lines, a pan caspase inhibition assay was

performed to determine if phenoxodiol functioned through manipulation of the intrinsic or

extrinsic caspase signalling pathways. Following treatment with both phenoxodiol and a pan-

caspase inhibitor, an MTS proliferation assay measuring cell proliferation was performed.

Figure 40 demonstrates the effect of a 10µM treatment of pan-caspase inhibitor Z-VAD-FMK

on the proliferation of cells in conjunction with 30µM of phenoxodiol after 48 hours. None of

the cell lines exhibited a significant difference between their respective DMSO vehicle control

group and 10µM caspase inhibitor (CI) only treatments, nor the 30µM PXD and combined

10µM CI 30µM PXD treatment groups. The LNCaP cell 30µM PXD only treatment group was

found to be significantly decreased in proliferation versus DMSO vehicle control (p=0.048) and

the 10µM CI only treatment (p≤0.001). The LNCaP cell combined 10µM CI 30µM PXD

treatment had the same result with a significant decrease in cell proliferation versus DMSO

vehicle control (p=0.0024) and 10µM CI only treatments (p≤0.001). DU145 cells did not follow

the same trend strictly, the 30µM PXD (p<0.05) and combined 10µM CI 30µM PXD (p=0.029)

groups were significantly different versus DMSO vehicle control but not statistically different

to the 10µM CI only treatment. PC3 cells exhibited the same response to caspase inhibition

and phenoxodiol treatment as LNCaP cells. The PC3 cells 30µM PXD group was significantly

different to DMSO vehicle control (p≤0.001) and the 10µM CI only (p=0.0062) groups. The PC3

cells combined 10µM CI 30µM PXD group was significantly different to DMSO vehicle control

(p≤0.001) and the 10µM CI only (p=0.0058) groups.

Page 183: Investigating the effects of, the isoflavone, Phenoxodiol

163

FIGURE 40 CASPASE INHIBITION TREATMENT WITH 10µM Z-VAD-FMK (CI) AND 30µM PHENOXODIOL

(PXD) OVER 48 HOURS

* indicates significance relative to control for that time point. ** indicates significance relative

to 10µM CI treatment and Control for that time point.

0

20

40

60

80

100

120

48 Hours

% C

ell

Gro

wth

vs

Co

ntr

ol

Treatment

LNCaP Z-VAD-FMK Caspase Inhibition

Negative

Control

CI 10µm

PXD 30µm

CI + PXD

** **

0

20

40

60

80

100

120

48 Hours

% C

ell

Gro

wth

vs

Co

ntr

ol

Treatment

DU145 Z-VAD-FMK Caspase Inhibition

Negative

Control

CI 10µm

PXD 30µm

CI + PXD

* *

0

20

40

60

80

100

120

48 Hours

% C

ell

Gro

wth

vs

Co

ntr

ol

Treatment

PC3 Z-VAD-FMK Caspase Inhibition

Negative

Control

CI 10µm

PXD 30µm

CI + PXD

** **

Page 184: Investigating the effects of, the isoflavone, Phenoxodiol

164

Chemotherapy and radiation treatments are known to induce oxidative stress in cancer cells

by rapid accumulation of highly reactive molecules such as nitric oxide (NO). After determining

that pan-caspase inhibition did not affect phenoxodiols ability to prevent the proliferation of

prostate cancer cells, a nitric oxide measurement assay, the Griess assay, was performed to

determine if phenoxodiol induced the production of nitric oxide, a potent free radical which

can induce cell cytotoxicity through oxidative stress. When this occurs it results in damage to

cell structures and potentially activates the apoptotic cascade, making it a target for

chemotherapy agents (Dharmarajan et al. 1999). Figure 41 demonstrates the amount of nitric

oxide produced by phenoxodiol addition to prostate cancer cells over 24 and 48 hours and

includes the positive control nitric oxide stimulators DEAN and SNP. None of the cell lines

exhibited significant changes in nitric oxide levels in comparison to DMSO vehicle control at

either 10µM or 30µM concentrations of phenoxodiol after 24 or 48 hours of treatment. LNCaP

cells exhibited a 1000% increase in NO production versus DMSO vehicle control under the

influence of DEAN (p<0.001) and while SNP effectively quadrupled control (p<0.001) it was

dwarfed by DEAN’s over 1000% increase in NO production and was significantly lower

(p<0.001) than DEAN stimulated production. DU145 and PC3 cell lines displayed a similar result

with a significant increase in NO production over DMSO vehicle control by DEAN stimulated

cells (DU145 p<0.001, PC3 p<0.001). In both PC3 and DU145 cell lines SNP stimulated an

increased production of NO versus DMSO vehicle control (DU145 p<0.001, PC3 p<0.001) and

continued the trend of SNP stimulated cells responding much lower than DEAN stimulated

cells (DU145 p<0.001, PC3 p<0.001). Both DU145 and PC3 cell lines had 1000% increase in NO

production under the influence of DEAN, similar to LNCaP; however, the SNP production was

only a 200% increase in NO production versus DMSO vehicle control whereas the LNCaP cells

had a 400% increase. This suggests the late stage representative cell lines, DU145 and PC3,

were less responsive to SNP stimulation than the early stage representative, LNCaP.

Phenoxodiol does not induce the production of reactive oxygen species in prostate cancer

cells.

Page 185: Investigating the effects of, the isoflavone, Phenoxodiol

165

FIGURE 41 NITRIC OXIDE FORMATION IN PROSTATE CANCER CELLS OVER 24 AND 48 HOURS POST

PHENOXODIOL TREATMENT MEASURED VIA GRIESS ASSAY

0

400

800

1200

24 Hours 48 Hours% N

O P

rod

uct

ion

vs

Co

ntr

ol

Treatment

LNCaP Nitric Oxide Production Following Phenoxodiol Treatment

Negative

Control

10µM PXD

30µM PXD

DEAN

SNP

*

**

0

400

800

1200

24 Hours 48 Hours% N

O P

rod

uct

ion

vs

Co

ntr

ol

Treatment

DU145 Nitric Oxide Production Following Phenoxodiol Treatment

Negative

Control

10µM PXD

30µM PXD

DEAN

SNP

0

400

800

1200

24 Hours 48 Hours% N

O P

rod

uct

ion

vs

Co

ntr

ol

Treatment

PC3 Nitric Oxide Production Following Phenoxodiol Treatment

Negative

Control

10µM PXD

30µM PXD

DEAN

SNP

**

Page 186: Investigating the effects of, the isoflavone, Phenoxodiol

166

* indicates significance relative to control for that time point. ** indicates significance relative

to 10µM DEAN treatment and control for that time point.

5.5. DISCUSSION

Two of the hallmarks of cancer initiation and progression are the ability of cells to become

insensitive to anti-growth signals and to evade apoptosis. To do this the cells accumulate

increased expression of anti-apoptotic signalling molecules, which prevents activation of the

intrinsic and extrinsic signalling pathways that normally induce cell death (Hanahan and

Weinberg 2000). Here we report that phenoxodiol does not induce cytotoxicity through direct

manipulation of the expression of key elements of the extrinsic and intrinsic cell death

pathways. In the previous chapter we characterised the prostate cancer cells’ response to

phenoxodiol’s treatment and determined that cytotoxicity appears to be partly caused by

significant mitochondrial depolarisation after treatment. This results in a significant decrease

in cell population and increase in cell death via apoptosis and late stage apoptosis. The JC-1

study gave an indication that mitochondrial catastrophe may be occurring, which can affect

AIF, Bax, Bcl-xL, Caspase-3 and xIAP expression.

A key aspect of chemotherapeutic treatments is the ability to induce cytotoxicity in multiple

stages of tumour development. Treatments with a known, consistent method of action

maximise potential treatment candidates and have increased success rates. Successful

treatments, which induce cytotoxicity and target the extrinsic or intrinsic apoptotic signalling

pathways, are focussed on due to the well characterised ability of these pathways to induce

inflammation-free cell death (Brown and Wouters 1999; Fojo and Bates 2003). To elucidate the

mechanism by which phenoxodiol elicits an anti-proliferative effect in prostate cancer cell

lines, we investigated the kinetics of apoptosis induction. In the previous chapter we reported

Page 187: Investigating the effects of, the isoflavone, Phenoxodiol

167

a lack of Caspase-3 activation, or decrease, within the first 48 hours for LNCaP and PC3 cell

lines, with DU145 having a subsequent increase at the 48 hour time point. This was surprising

because we detected significant levels of apoptosis in LNCaP and DU145 cell lines with PC3

cells responding necrotically to treatment, and other studies have indicated the ability of

phenoxodiol to induce apoptosis through PARP-1 activation (Aguero et al. 2010). The

significant increase in Caspase-3 expression in DU145 cells, although minor, was in response to

an increase in Caspase-3 signalling following 48 hours of treatment. This confirmed that, in

some prostate cancer cell lines, phenoxodiol can induce a downstream classic apoptotic

response as previously reported (Kamsteeg et al. 2003; Aguero et al. 2005; Alvero et al. 2006;

Yu et al. 2006).

Prostate cancer has been shown to express Bcl-2 family members (Scherr et al. 1999) as well as

having changes in the ratio of pro- and anti-apoptotic signalling molecules following treatment

(Tang et al. 2006). However, a lack of change has also been previously reported in clonogenic

studies (Tannock and Lee 2001). The over expression of Bcl-xL leads to resistance to

chemotherapeutic drugs and radiotherapy, while depletion of Bcl-xL by shRNA, antisense

cDNA, or antisense oligodeoxynucleotides can impair tumour cell growth, sensitise cells to

apoptosis inducing agents, and increase cell sensitivity to chemotherapy (Wang et al. 2008).

Members of the BH3 family, such as Bcl-2 and Bcl-xL, are anti-apoptotic and maintain

mitochondrial membrane integrity by binding mitochondrial porin channels. These anti-

apoptotic proteins also protect cells from apoptosis by binding to pro-apoptotic BH3-only

members, such as Bid and Bax (Korsmeyer et al. 2000; Cheng et al. 2001; Cory and Adams

2002; Puthalakath and Strasser 2002). The ratio of pro-apoptotic Bax protein to anti-apoptotic

Bcl-xL can be a determinant of cellular susceptibility to apoptosis. The release of Bax from the

mitochondria in LNCaP, DU145 and PC3 cells was exhibited in the previously reported JC-1

mitochondrial depolarisation assay.

Page 188: Investigating the effects of, the isoflavone, Phenoxodiol

168

In this study we determined that the cytotoxicity of phenoxodiol was not associated with

consistent changes in both pro-apoptotic and anti-apoptotic members of the Bcl-2 family

between the cell lines, contrasting previous studies (Yu et al. 2006). LNCaP displayed a

decreased expression of Bax protein and increased expression of Bcl-xL protein, with cells

attempting to prevent the effect of phenoxodiol induced mitochondrial depolarisation by

mediating the effects of the Bcl-2 family. However, LNCaP cells still exhibited significant

cytotoxicity to treatment, indicating that the classic intrinsic apoptotic pathway might not be

directly responsible for the effect of phenoxodiol treatment. DU145 cells did not express Bax

protein at all, influencing the cell into a strong anti-apoptotic phenotype: yet a detectable

decrease in Bcl-xL protein was determined after treatment, indicating a susceptibility to

phenoxodiol induced mitochondrial depolarisation with resultant mitochondrial efflux. PC3

cells levels of Bax were unchanged following treatment, but the expression of Bcl-xL levels

were significantly reduced, indicating that the ratio of Bcl-xL:Bax had significantly decreased, a

sign of cell death signalling sensitivity from this notoriously chemoresistant cell line (Li et al.

2008).

Modulation of other pro- or anti-apoptotic proteins also remained inconsistent, indicating that

phenoxodiol could be inducing cytotoxicity through means other than the classical intrinsic and

extrinsic cell death pathways, with DU145 cells exhibiting downstream effects of these other

pathways. Phenoxodiol mediated cell death was not linked to the Caspase-independent

pathway of Apoptosis Inducing Factor (AIF). AIF induces caspase independent cell death

primarily through translocation from the mitochondria to the nucleus, where it mediates

chromatin condensation and high-molecular-weight fragmentation of DNA (Hong et al. 2004).

AIF expression was only mediated in the LNCaP cell line, with DU145 not expressing the

protein. Expression in LNCaP cells was found to decrease in response to decreased mRNA

Page 189: Investigating the effects of, the isoflavone, Phenoxodiol

169

signalling, which is counter intuitive to the effects of phenoxodiol as it was expected to

increase this expression to induce the cytotoxic effects that were exhibited. PC3 cells were not

determined to have much AIF signalling alteration except for a small decrease after 24 hours of

treatment, while the 48 hour treatment group remained unchanged.

The lack of a consistent change in gene expression between LNCaP, DU145 and PC3 cell lines

indicates that phenoxodiol does not promote cell death by disrupting the stoichiometry of key

pro- and anti-apoptotic molecules investigated here. The common model for cancer therapy

activity has been that anti-cancer regimens cause apoptotic signalling to occur, often mediated

by p53 and Caspase-3 upregulation (Fojo and Bates 2003). Therefore, cells that are resistant to

apoptosis achieve this through forced over expression of anti-apoptotic molecules such as xIAP

and Bcl-xL family members or p53 mutation (Brown and Attardi 2005). Members of the

Inhibitors of Apoptosis Proteins (IAP), such as xIAP, can prevent apoptosis by inactivating

Caspase-3 and Caspase-9. The results reported in this study are in contrast to other reports,

where it has been demonstrated in ovarian cancer that caspase activation, Bax upregulation

and xIAP degradation are key proteins modulated by phenoxodiol (Sapi et al. 2004; Alvero et

al. 2006; Mor et al. 2006; Kluger et al. 2007). We have previously established a link between

chemoresistance and xIAP expression in refractory ovarian cancer (Amantana et al. 2004). We

detected only decreased xIAP expression in the DU145 cells following 24 hours of treatment.

This decrease coincides with an increase in Caspase-3 mRNA signalling and activated

expression, reinforcing the ability of phenoxodiol to induce DU145 cells into an intrinsically

mediated apoptotic response downstream of the pathway phenoxodiol is interacting with.

The inability of the pan caspase inhibitor, Z-VAD-FMK, to prevent the anti-proliferative activity

of phenoxodiol in all three cell lines reinforces that the phenoxodiol induced mechanism of cell

Page 190: Investigating the effects of, the isoflavone, Phenoxodiol

170

death in prostate cancer is occurring via a caspase-independent pathway. This is in contrast to

many studies that indicate PARP and caspase activation as the effect of phenoxodiol treatment

(Kamsteeg et al. 2003; Aguero et al. 2005). One potential method of induction investigated

was the ability of phenoxodiol to induce accumulation/production of reactive oxygen species

(ROS) in the cell, causing oxidative stress. Mitochondrial swelling and volume increase due to

depolarisation is linked to ROS production, where oxidative stress induces the expression of

specific genes that contain an antioxidant response element (ARE) in their promoters. ARE

promoted genes can trigger apoptosis or necrosis of various cell types and, in several

instances, can inhibit cell growth and interfere with the cell cycle (Juhaszova et al. 2004;

Minelli et al. 2009). The cell cycle modulator p21WAF1 has been implicated in protecting the cell

from NO induced cell death (Yang et al. 2000) and it has also been reported that anti-apoptotic

Bcl-2 family members, such as Bcl-xL, reduce accumulation of reactive oxygen species (ROS) in

transfected cells and can protect a variety of cells from apoptosis induced by oxidative

stressors. Conversely, Bax expression causes increased mitochondrial production of ROS and

apoptosis when over expressed in cultured cells (Dharmarajan et al. 1999). This study

investigated the effects of phenoxodiol in producing nitric oxide (NO), a potent ROS molecule,

after the effect of phenoxodiol on mitochondrial depolarisation was exhibited. We discovered

that, in all cell lines, phenoxodiol did not induce NO production as a response to treatment,

but that all three cell lines retained the ability to have NO production stimulated by DEAN and

SNP addition.

Recent reports indicate that, in addition to apoptosis, cancer cells can be effectively

eliminated through necrosis, mitotic catastrophe and premature senescence, which result in

cell cycle arrest and subsequent cell death signalling (Brown and Wouters 1999; Brown and

Wilson 2003; Tannock et al. 2004). Cells that are non-responsive to treatment through the

classic intrinsic apoptosis pathways, yet have an apoptotic response, must activate this process

Page 191: Investigating the effects of, the isoflavone, Phenoxodiol

171

through other mechanisms such as mitochondrial response and depolarisation, mitotic

catastrophe and DNA degradation (Ruth and Roninson 2000; Brown and Wilson 2003).

Additionally, cells can undergo apoptosis through mechanisms such as Cytochrome c release,

Caspase-8 and Caspase-9 signalling, as well as high-molecular-weight DNA breakdown. We

previously exhibited the ability of phenoxodiol induction to induce efflux of mitochondria and,

therefore, Cytochrome c from the mitochondria into the cytoplasm. Indeed, it would appear

from our studies of cell lines representative of different phases of prostate cancer, that

phenoxodiol elicits an alternative induction of cytotoxicity instead of classical extrinsic and

intrinsic apoptotic signalling.

This study indicates that phenoxodiol exhibits a significant ability to induce cytotoxicity and cell

death in the prostate cancer cell lines LNCaP, DU145 and PC3 through different signalling

pathways than those previously reported in ovarian cancer studies, and clearly demonstrates

the caspase independent induction of this cell death. This study also determined that nitric

oxide production and, therefore, oxidative stress is not altered after treatment. Coupled with

the induction of apoptotic cell death through non-classical pathways, phenoxodiol shows

potential as a drug for future treatment of prostate cancer that is resistant to apoptotic

signalling.

Page 192: Investigating the effects of, the isoflavone, Phenoxodiol

172

6 . C H A P TER S IX : TH E E F F EC TS OF P H ENOX O D IOL ON T H E C EL L C YC L E

6.1. INTRODUCTION

Two of the hallmarks of cancer are the ability to be self-sufficient in growth signals and to have

infinite replicative potential, which can be initiated through damaging the cell cycle restriction

points, thereby allowing for progression of cancer from benign to metastatic states (Hanahan

and Weinberg 2000). Dysregulated growth rates can initiate cancer as more cell cycle

machinery becomes damaged, mutations are allowed to accumulate, methylation status

changes occur and cells become undifferentiated and progress towards metastasis. The cell

cycle is composed of G1, S, G2 and M phases, which represent normal function, DNA

replication, organelle replication and mitotic separation respectively (Marieb 2012). Any errors

that may occur at these steps could be catastrophic for normal cell functioning and, as such,

the cell cycle apparatus retains potent signalling molecules that can search for errors and

rapidly induce apoptosis (Cooper 2003). The initial point where this process can occur is

referred to as the G1 restriction point (Pardee 1974).

To maintain normal function, the cell cycle is a potent inducer of cell death through regulation

of DNA content at a variety of restriction points. As the cell reaches a restriction point, the

gene p53 detects and attempts to repair any evident DNA damage; if the damage is too great it

can initiate apoptosis. The p53 gene is a potent tumour suppressor that exerts its functions

through activation of downstream targets, some of which include induction of CDKN1/p21WAF1,

14-3-3y, and REPRIMO for G1-G2 arrest; p53R2 for DNA repair and Bax, Puma, p53AIP1, PERP,

and CD95 for apoptosis (Wasylishen and Penn 2010). Unfortunately p53 is found mutated in

nearly 70% of carcinomas. Considerable evidence indicates that the choice made by p53 to

Page 193: Investigating the effects of, the isoflavone, Phenoxodiol

173

activate cell cycle arrest and DNA repair pathways, or the apoptosis pathway after DNA

damage, is dependent on the extent of unrepaired or misrepaired double-strand breaks in the

DNA (Devlin et al. 2008). Disabling the p53 pathway enables cells to enter and proceed

through the cell cycle under conditions that increase the frequencies of aneuploidy, gene

amplification, deletion and translocation. This, coupled with loss of p53 dependant apoptosis,

increases genetic instability and is highly selected during cancer progression (Vafa et al. 2002).

Cell cycle progression is intricately regulated by the interactions of cyclin and cyclin-dependent

kinase (Cdk) complexes. Cyclin-D1 is a regulatory subunit of the highly conserved cyclin family

that phosphorylates and, together with sequential phosphorylation by Cyclin-E/CDK2,

inactivates the cell-cycle inhibiting function of the retinoblastoma protein. Upon mitogenic

stimulation, the Cyclin-D Cdk4/Cdk6 and Cyclin-E Cdk2 complexes mediate phosphorylation of

retinoblastoma protein, which induces transcriptionally active E2F thereby ensuring G1-S

transit (Baldin et al. 1993; Roy et al. 2007). Retinoblastoma protein serves as a gatekeeper of

the G1 phase, and passage through the restriction point leads to DNA synthesis, making Cyclin-

D1 expression an important target to regulate cancer cell replication (Fu et al. 2004).

P21WAF1 is a cyclin dependant kinase inhibitor (CdkI) family member, along with p27 and p57,

which interfere with the cyclin dependant kinase-cyclin complex. The group is regulated both

by internal and external signals, with the expression of p21WAF1 under transcriptional control of

the p53 tumour suppressor gene (Srivastava et al. 2007). As an important mediator of

apoptosis, the frequency of somatic mutations in the p21WAF1 gene and family in cancers is very

rare; which underlines the importance of these molecules as promising therapeutic targets.

Both p21 and p27 are upregulated by a variety of regulatory pathways at the transcriptional as

well as post-transcriptional levels, with p21WAF1 transcriptionally upregulated by p53 in

Page 194: Investigating the effects of, the isoflavone, Phenoxodiol

174

response to DNA damage. p21WAF1 is also activated by various transcription factors and

subsequently mediates growth arrest, senescence and apoptosis in a p53-independent

manner, and p21 mRNA stability can be post transcriptionally regulated by HuR, an RNA-

binding protein, in response to stress (Roy et al. 2007; Roy et al. 2008).

An important indicator of cell proliferation rate, Ki-67 is present in all proliferating cells,

normal and neoplastic, and its level of expression indicates the rate of growth of a cell

population. Studies proved the hypothesis that, in breast cancer, the Ki-67 index is an

independent prognostic factor in both patient survival assessment and disease recurrence. The

Ki-67 index can also be used as a predictive factor of neoplastic cell response to certain types

of therapy as well as a measure capable of indicating success of treatment (Koda et al. 2007).

C-Myc is an important regulator of cell function that was implicated as one of the first

oncogenes. It is known that cell cycle regulation is altered under excess c-Myc expression with

a decrease in time taken to reach the restriction point of G1. Cell cycle regulation has been

implicated by varied means; canonical Wnt/β-catenin stimulation, Cyclin-D and Cyclin-E CDK2

stimulation, inhibiting the effect of p27 a CdkI family member amongst others (Mateyak et al.

1999). Conversely, c-Myc expression has also been promoted as a control mechanism for cell

cycle-induced apoptosis and prevention of differentiation in normal functioning cells

(Wasylishen and Penn 2010).

Initiating the normal and cancerous prostate cell cycle progression is the activity of the growth

factor androgen receptor (AR), which is found to be mutated in most late stage prostate

cancers. The AR is widely accepted to be the initiator of prostate cell replication and

differentiation signalling through transcription of mitogenic and anti-apoptotic factors. Cell

signalling pathways, such as the canonical Wnt/β-catenin signalling pathway, drive the cells

Page 195: Investigating the effects of, the isoflavone, Phenoxodiol

175

towards growth and differentiation while interaction with anti-apoptotic pathways, such as the

Bcl-xL pathway, prevents the cells from entering into apoptosis when the AR is stimulated (Sun

et al. 2008; Wang et al. 2008). Because AR signalling has also been shown to play a key role in

prostate carcinogenesis, androgen ablation therapy is a commonly used form of treatment,

particularly for advanced disease. While ablation therapy leads to initially significant levels of

prostate cancer cell apoptosis, the effect is short-lived and ultimately not curative because

most patients develop androgen-independent/hormone refractory disease (D'Antonio et al.

2008). Manipulation of this growth factor by upregulation of receptors, mutation to a

constitutively active form or a complete loss of AR as a mitogenic requirement are

characteristics of late stage prostate cancer (Peng et al. 2008).

Levels of β-catenin and AR are both increased in hormone refractory prostate cancer, β-

catenin is activated by canonical Wnt stimulation, but mutated forms of β-catenin, which can

result in a permanently stabilised protein, have also been detected in prostate cancer. β-

catenin has multiple functions that involve both cell adhesion and signal transduction in

response to Wnt ligands (Chesire et al. 2002). In the absence of Wnt ligands, GSK3β complexes

with other proteins and degrades cytoplasmic β-catenin, but when Wnt ligands bind to the

frizzled receptor complex, the resulting activation of the cytoplasmic protein dishevelled

inactivates GSK3β, thereby preventing degradation of β-catenin (Wan et al. 2012). It is

important to note that GSK3β has been linked to prevention of AR transcriptional activity; so

Wnt ligand binding both increases AR expression and β-catenin stability (Li et al. 2008). Upon

stimulation, β-catenin translocates to the nucleus where it can drive cell cycle progression

through interaction with T-cell factor (TCF)/lymphoid enhancer factor (LEF) transcription

factors to initiate transcription of target genes such as c-Myc and Cyclin-D1, driving cell

growth. As such, dysregulation of the Wnt, AR or β-catenin pathways can result in a

metastatic, highly replicative phenotype (Wang et al. 2008). It has been postulated that the

Page 196: Investigating the effects of, the isoflavone, Phenoxodiol

176

ratio of β-catenin/AR might be an important prognostic indicator that may even help define a

subpopulation of men with prostate cancer for individualised management (Wan et al. 2012).

Secreted frizzled related protein 4 (sFRP4) is one of a group of proteins that antagonise the

canonical Wnt/β-catenin pathway, decreasing Wnt activity and preventing activated β-catenin

from forming. sFRP4 can also decrease invasiveness in androgen-independent prostate cancer

cells and has been shown to be anti-angiogenic and pro-apoptotic in nature (Muley et al.

2010). Moreover, the correlation between increased membranous sFRP4 and β-catenin

expression in a large human cohort supports evidence for sFRP4 as a prognostic marker in

localised androgen-dependent prostate cancer. Unlike other inhibitors of Wnt signalling, sFRP4

appears to affect androgen-dependent and androgen-independent prostate cancer (Horvath et

al. 2004; Horvath et al. 2007).

In this chapter we explored the ability of phenoxodiol to impact the cell cycle of prostate

cancer cells and investigate the underlying signalling pathways c-Myc, Cyclin-D1, Ki-67 and

p21WAF1. We then investigated whether phenoxodiol alters the expression of the active protein

form of β-catenin, which is known to be important in prostate homeostasis and whose

expression has been implicated in poor prognosis. Finally, we investigated sFRP4 protein

expression, after phenoxodiol treatment because sFRP4 is an inhibitor of the canonical Wnt/β-

catenin signalling pathway and a potential inhibitor of prostate cancer cell growth.

Page 197: Investigating the effects of, the isoflavone, Phenoxodiol

177

6.2. AIMS

The aims of this chapter were as follows:

Aim 1: To determine if phenoxodiol affected the cell cycle phases of the prostate cancer lines

LNCaP, DU145 and PC3.

Aim 2: To determine if signalling pathways c-Myc, Cyclin-D1, Ki-67, p21WAF1 were altered post

phenoxodiol treatment.

Aim 3: To determine if the expression of sFRP4 ligand was altered in response to phenoxodiol

treatment.

6.3. METHODOLOGY

The methodology utilised in this chapter is discussed in detail in the materials and methods

chapter (page 50). Analysing the population of cells in each cell cycle phase indicated whether

phenoxodiol had a direct effect upon the cell cycle of prostate cancer cells. Briefly, cells were

seeded into 6 well plates and incubated for 48 hours before 24 or 48 hours of treatment media

were applied. Cell media were aspirated and centrifuged and cells were trypsinised, then

placed into the same tube and centrifuged. Cells were washed with PBS and then -20°C 70%

ethanol was added drop wise, while vortexing to fixate the cells. Cells were stored at 4°C then

rehydrated with PBS and stained with propidium iodide before being placed into a FACS Canto

II cytometer, which determined fluorescence and therefore DNA content. FlowJo software was

utilised to further analyse the sample and distinguish quantitative populations of G1, S and G2

phase cells.

Page 198: Investigating the effects of, the isoflavone, Phenoxodiol

178

Once it was determined that phenoxodiol had an impact on the cell cycle it was necessary to

explore the underlying cell cycle and cell proliferation signalling that could be the target of

phenoxodiol treatment. Quantitative PCR (qPCR) analysis was performed on all three cell lines,

LNCaP, DU145 and PC3 after 24 and 48 hours of 10µM and 30µM phenoxodiol treatment. As

previously states, cells were seeded, treated and then RNA extraction was performed using the

Chomczynski guanidinium thiocyanate-phenol-chloroform method (Chomczynski and Sacchi

1987). Purified RNA was analysed and converted into cDNA then cleaned with a post PCR kit.

Each qPCR experiment was performed using 2µL of cDNA per sample and primers were

designed to test for c-Myc, Cyclin-D1, Ki-67 and p21WAF1 by using published primers. Primer

spanning of an intron/exon boundary was confirmed using Primer3 and BLAST search engines.

Quantitative PCR (qPCR) was performed as previously states, using 2µL of cDNA to a master

mix of primers/Taq and then applying 35 temperature cycles at varying annealing

temperatures on a Corbett Rotorgene 3000 or 6000. The standards used were 10 fold dilutions

of a purified gel extract and were given concentrations. The resulting standard curve then

determined final sample concentrations compared against the expression of L19. Protein level

analysis was performed using a Western blot protocol as previously described. Cells were

seeded, treated and protein extracted with a RIPA buffer/β-mercaptoethanol based, whole cell

lysate method. After protein concentration was determined it was loaded onto a SDS-

Page/acrylamide based gel and current applied to the gel. After protein separation, transfer to

a nitrocellulose membrane was performed and Ponceau red used to determine effective

transfer. Finally, primary antibodies were added to the membrane and incubated in the

appropriate conditions, excess was removed with TBS-T washes and a secondary antibody

conjugated to HRP applied in the appropriate conditions. A chemiluminescent substrate kit

was then used to detect antibody binding concentration and results compared to β-actin

protein.

Page 199: Investigating the effects of, the isoflavone, Phenoxodiol

179

6.4. RESULTS

In the previous chapter we determined that apoptotic signalling was not consistently altered in

response to phenoxodiol treatment and, therefore, it was not directly targeting one specific

pathway tested. Another of the hallmarks of cancer, the ability to limitlessly replicate, was

investigated through looking at the cell cycle response to phenoxodiol treatment. After 24 or

48 hours of phenoxodiol treatment, cells were stained with propidium iodide and analysed to

determine DNA content, which was then assessed into cell cycle phase populations G1, S and

G2 phase using FlowJo software.

Figure 42 demonstrates the LNCaP cell line cell cycle response to 10µM and 30µM phenoxodiol

treatment over 24 and 48 hours by assessing the cell cycle phase populations differentiated by

DNA content. Phenoxodiol induced significantly decreased G2 phase cell populations versus

DMSO vehicle control, over 24 hours for both 10µM (p<0.001) and 30µM (p<0.001)

phenoxodiol treatments in LNCaP cells. The S phase cell population was found to increase

versus DMSO vehicle control following 24 hours of 10µM (p<0.0021) and 30µM (p=0.0016)

phenoxodiol treatment. The 10µM and 30µM phenoxodiol treatment groups were not

significantly different versus each other after 24 hours.

Phenoxodiol induced significantly decreased G2 phase cell populations over 48 hours, versus

DMSO vehicle control for both 10µM (p<0.001) and 30µM (p=0.0016) treatments in LNCaP

cells. Only the 10µM phenoxodiol treatment was determined to increase the S phase cell

population (p=0.0056) after 48 hours but the 30µM phenoxodiol treatment had a significantly

increased cell population in G1 phase versus both DMSO vehicle control (p=0.0013). In LNCaP

cells phenoxodiol was determined to consistently decrease G2 phase cell population over 24

and 48 hours following treatment with 10µM and 30µM concentrations.

Page 200: Investigating the effects of, the isoflavone, Phenoxodiol

180

FIGURE 42 LNCAP CELL CYCLE ANALYSIS AFTER 24 AND 48 HOURS OF 10µM AND 30µM PHENOXODIOL

TREATMENT

* indicates significance relative to control for that time point. ** indicates significance relative

to 10µM Phenoxodiol treatment and control for that time point.

Figure 43 demonstrates the DU145 cell line cell cycle response to 10µM and 30µM

phenoxodiol treatment over 24 and 48 hours by assessing the cell cycle phase populations

differentiated by DNA content. Following 24 hours of phenoxodiol treatment the DU145 G2

phase cell population was significantly decreased versus DMSO vehicle control in both 10µM

(p=0.0028) and 30µM (p=0.0021) concentrations. Only the 10µM phenoxodiol treatment had

an increased S phase cell population versus DMSO vehicle control (p<0.001) but the 30µM

0

20

40

60

80

100

G1 S G2

% o

f C

ell

Po

pu

lati

on

Treatment

LNCaP Cell Cycle Response to 24 Hours of PXD Treatment

Control

10µM PXD

30µM PXD

0

20

40

60

80

100

G1 S G2

% o

f C

ell

Po

pu

lati

on

Treatment

LNCaP Cell Cycle Response to 48 Hours of PXD Treatment

Control

10µM PXD

30µM PXD

Page 201: Investigating the effects of, the isoflavone, Phenoxodiol

181

treatment had an increased G1 phase cell population versus DMSO vehicle control (p=0.0038)

and 10µM (p=0.0049).

Phenoxodiol induced significantly decreased G2 phase cell populations in DU145 cells, over 48

hours for both 10µM (p<0.001) and 30µM (P<0.001) treatments versus DMSO vehicle control

and between the treatment concentrations (P<0.001). The S phase cell population was

significantly increased versus DMSO vehicle control in both 10µM (p<0.001) and 30µM

(p<0.001) treatments as well as significantly different between treatments (p<0.001). The G1

phase cell population was significantly decreased versus DMSO vehicle control for the 10µM

treatment (p<0.001) and significantly increased versus DMSO vehicle control (p<0.001) and

versus 10µM (p<0.001) for the 30µM treatment group. In DU145 cells phenoxodiol was

determined to consistently decrease G2 phase cell population over 24 and 48 hours following

treatment with 10µM and 30µM concentrations.

Page 202: Investigating the effects of, the isoflavone, Phenoxodiol

182

FIGURE 43 DU145 CYCLE ANALYSIS AFTER 24 AND 48 HOURS OF 10µM AND 30µM PHENOXODIOL

TREATMENT

* indicates significance relative to control for that time point. ** indicates significance relative

to 10µM Phenoxodiol treatment and control for that time point.

0

20

40

60

80

100

G1 S G2

% o

f C

ell

Po

pu

lati

on

Treatment

DU145 Cell Cycle Response to 24 Hours of PXD Treatment

Control

10µM PXD

30µM PXD

0

20

40

60

80

100

G1 S G2

% o

f C

ell

Po

pu

lati

on

Treatment

DU145 Cell Cycle Response to 48 Hours of PXD Treatment

Control

10µM PXD

30µM PXD

Page 203: Investigating the effects of, the isoflavone, Phenoxodiol

183

Figure 44 demonstrates the PC3 cell line cell cycle response to 10µM and 30µM phenoxodiol

treatment over 24 and 48 hours by assessing the cell cycle phase populations differentiated by

DNA content. Following 24 hours of phenoxodiol treatment the PC3 G2 phase cell population

was significantly decreased versus DMSO vehicle control in both 10µM (p<0.001) and 30µM

(p=0.0071) concentrations. No significant differences were detected in G1 phase cell

population but S phase cell populations were significantly increased versus DMSO vehicle

control for 10µM (p<0.001) and 30µM (p=0.0022) treatments.

Following 48 hours of treatment the PC3 G2 phase cell population was significantly decreased

versus DMSO vehicle control in both 10µM (p<0.001) and 30µM (p<0.001) treatments as well

as significantly different between the treatments (p<0.001). The PC3 S phase cell population

was significantly increased versus DMSO vehicle control in both 10µM (p<0.001) and 30µM

(p<0.001) and significantly different between treatments (p<0.001). The PC3 G1 phase cell

population was significantly decreased versus DMSO vehicle control in both 10µM (p<0.001)

and 30µM (p<0.022) treatments and significantly different between the treatments (p<0.001).

In PC3 cells phenoxodiol was determined to consistently decrease G2 phase cell population

over 24 and 48 hours following treatment with 10µM and 30µM concentrations.

Page 204: Investigating the effects of, the isoflavone, Phenoxodiol

184

FIGURE 44 PC3 CELL CYCLE ANALYSIS AFTER 24 AND 48 HOURS OF 10µM AND 30µM PHENOXODIOL

TREATMENT.

* indicates significance relative to control for that time point. ** indicates significance relative

to 10µM Phenoxodiol treatment and control for that time point.

0

20

40

60

80

100

G1 S G2

% o

f C

ell

Po

pu

lati

on

Treatment

PC3 Cell Cycle Response to 24 Hours of PXD Treatment

Control

10µM PXD

30µM PXD

0

20

40

60

80

100

G1 S G2

% o

f C

ell

Po

pu

lati

on

Treatment

PC3 Cell Cycle Response to 48 Hours of PXD Treatment

Control

10µM PXD

30µM PXD

**

Page 205: Investigating the effects of, the isoflavone, Phenoxodiol

185

Once it was determined that phenoxodiol had an impact on the cell cycle, it was necessary to

explore the underlying cell cycle and cell proliferation signalling that could be the target of

phenoxodiol treatment. Quantitative PCR (qPCR) analysis was performed on all three cell lines,

with the housekeeping gene L19 used as a standardising agent, control expression was

designated as one, for ease of graphing.

c-Myc is a potent initiator of cell replication and has been implicated in increasing the rate at

which cells enter S phase (Wasylishen and Penn 2010). Figure 45 demonstrates the

quantitative mRNA expression of c-Myc in cells treated over 24 and 48 hour periods with

phenoxodiol. After 48 hours of 30µM phenoxodiol treatment, PC3 cells were found to

significantly increase the expression of c-Myc versus DMSO vehicle control (p=0.033). Neither

LNCaP nor DU145 cells were found to have any significant changes in c-Myc expression in

response to phenoxodiol treatment.

Page 206: Investigating the effects of, the isoflavone, Phenoxodiol

186

FIGURE 45 C-MYC MRNA EXPRESSION ANALYSIS OF PROSTATE CANCER OVER 24 AND 48 HOURS POST

PHENOXODIOL TREATMENT.

* indicates significance relative to control for that time point.

0

0.5

1

1.5

2

2.5

3

24 Hours 48 Hours

c-M

yc/L

19

Treatment

LNCaP c-Myc qPCR Expression

Control

10µM PXD

30µM PXD

0

0.5

1

1.5

2

2.5

3

24 Hours 48 Hours

c-M

yc/L

19

Treatment

DU145 c-Myc qPCR Expression

Control

10µM PXD

30µM PXD

0

0.5

1

1.5

2

2.5

24 Hours 48 Hours

c-M

yc/L

19

Treatment

PC3 c-Myc qPCR Expression

Control

10µM PXD

30µM PXD

Page 207: Investigating the effects of, the isoflavone, Phenoxodiol

187

Cyclin-D1 is recognised as potent initiator of cell cycle progression from G1 through to S phase

by the Cyclin-D1 Cdk4 complex activating the Cyclin E Cdk2 complex, which results in inhibition

of the cell cycle inhibiting Rb protein (Ladha et al. 1998). Figure 46 demonstrates the

quantitative mRNA expression of the cell cycle regulator gene Cyclin-D1 over 24 and 48 hours

post phenoxodiol treatment in prostate cancer cells. Decreasing Cyclin-D1 expression, in

response to phenoxodiol treatment, could result in quiescent and apoptotically sensitive cells.

LNCaP cells did not have a detectable change in Cyclin-D1 expression level under the influence

of phenoxodiol treatment. DU145 cells were found to have a significant decrease in the

expression of Cyclin-D1 versus DMSO vehicle control (p=0.0071) after 24 hours of treatment

with 30µM phenoxodiol no other changes in expression were detected in the DU145 cell line.

PC3 cells exhibited a similar trend to treatment as the DU145 cells, with both the 10µM

phenoxodiol (p=0.026) and 30µM phenoxodiol (p=0.0011) treatments significantly decreased

in expression versus DMSO vehicle control after a 24 hour period of treatment.

Page 208: Investigating the effects of, the isoflavone, Phenoxodiol

188

FIGURE 46 CYCLIN-D1 MRNA EXPRESSION ANALYSIS OF PROSTATE CANCER CELLS OVER 24 AND 48

HOURS POST PHENOXODIOL TREATMENT.

* indicates significance relative to control for that time point.

0

0.5

1

1.5

2

2.5

24 Hours 48 Hours

Cyc

linD

1/L

19

Treatment

LNCaP Cyclin-D1 qPCR Expression

Control

10µM PXD

30µM PXD

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

24 Hours 48 Hours

Cyc

linD

1/L

19

Treatment

DU145 Cyclin-D1 qPCR Expression

Control

10µM PXD

30µM PXD

0

0.2

0.4

0.6

0.8

1

1.2

24 Hours 48 Hours

Cyc

linD

1/L

19

Treatment

PC3 Cyclin-D1 qPCR Expression

Control

10µM PXD

30µM PXD

Page 209: Investigating the effects of, the isoflavone, Phenoxodiol

189

Ki-67 expression is an effective indicator of rate of proliferation in cells (Koda et al. 2007).

Figure 47 demonstrates the quantitative mRNA expression of the cell proliferation gene Ki-67

over 24 and 48 hours post phenoxodiol treatment in prostate cancer cells. LNCaP cells were

significantly decreased in Ki-67 mRNA signalling over 24 hours with both 10µM (p=0.0043) and

30µM (p=0.0065) phenoxodiol treatments exhibiting decreased mRNA expression versus

DMSO vehicle control. DU145 cells had no significant difference in Ki-67 mRNA expression over

24 and 48 hours of treatment with 10µM and 30µM PXD although a biological trend towards

decreased expression was indicated (p=0.082, p=0.1 respectively). PC3 cells exhibited a

significant decrease in Ki-67 mRNA signalling, over 24 hours, in both 10µM (p=0.034) and

30µM (p=0.036) phenoxodiol treatments while the 48 hour 10µM phenoxodiol treatment

exhibited significantly decreased mRNA expression versus DMSO vehicle control (p<0.05) and

30µM phenoxodiol (p=0.032).

Page 210: Investigating the effects of, the isoflavone, Phenoxodiol

190

FIGURE 47 KI-67 MRNA EXPRESSION ANALYSIS OF PROSTATE CANCER CELLS OVER 24 AND 48 HOURS

POST PHENOXODIOL TREATMENT.

* indicates significance relative to control for that time point. ** indicates significance relative

to 30µM Phenoxodiol treatment and control for that time point.

0

0.2

0.4

0.6

0.8

1

1.2

24 Hours 48 Hours

Ki-

67

/L1

9

Treatment

LNCaP Ki-67 qPCR Expression

Control

10µM PXD

30µM PXD

0

0.5

1

1.5

2

2.5

24 Hours 48 Hours

Ki-

67

/L1

9

Treatment

DU145 Ki-67 qPCR Expression

Control

10µM PXD

30µM PXD

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

24 Hours 48 Hours

Ki-

67

/L1

9

Treatment

PC3 Ki-67 qPCR Expression

Control

10µM PXD

30µM PXD

Page 211: Investigating the effects of, the isoflavone, Phenoxodiol

191

Figure 48 demonstrates the quantitative mRNA expression of the cell proliferation gene

p21WAF1 over 24 and 48 hours post phenoxodiol treatment in prostate cancer cells. P21WAF1 is a

cell cycle inhibiting factor that can prevent the formation of the Cyclin E Cdk2 complex. This

inhibition prevents the progression of the cell through G1 to S phase (Roy et al. 2007). LNCaP

cells exhibited a significant increase in p21 mRNA expression versus DMSO vehicle control over

24 hours of treatment with 10µM phenoxodiol (p=0.0099), and over 48 hours of treatment

with both 10µM (p<0.050) and 30µM phenoxodiol (p=0.011) concentrations. DU145 cells

exhibited a significant increase in p21 mRNA expression versus DMSO vehicle control after 48

hours of treatment with both 10µM (p=0.0048) and 30µM (p=0.028) phenoxodiol

concentrations. PC3 cells exhibited a significant increase in p21 mRNA expression over both 24

and 48 hours of treatment (p=0.042 and p=0.0044 respectively) with the 30µM phenoxodiol

treatment group.

Page 212: Investigating the effects of, the isoflavone, Phenoxodiol

192

FIGURE 48 P21 MRNA EXPRESSION ANALYSIS OF PROSTATE CANCER OVER 24 AND 48 HOURS POST

PHENOXODIOL TREATMENT.

* indicates significance relative to control for that time point.

0

2

4

6

8

10

12

24 Hours 48 Hours

p2

1/L

19

Treatment

LNCaP p21 qPCR Expression

Control

10µM PXD

30µM PXD

0

0.5

1

1.5

2

2.5

3

3.5

4

24 Hours 48 Hours

p2

1/L

19

Treatment

DU145 p21 qPCR Expression

Control

10µM PXD

30µM PXD

0

5

10

15

20

25

24 Hours 48 Hours

p2

1/L

19

Treatment

PC3 p21 qPCR Expression

Control

10µM PXD

30µM PXD

Page 213: Investigating the effects of, the isoflavone, Phenoxodiol

193

β-catenin is the activated downstream protein of the canonical Wnt/β-catenin signalling

pathway that can interact with androgen receptor and, as one from a list of potential targets,

activate Cyclin-D1 and therefore the cell cycle (Yu et al. 2011). Figure 49 demonstrates the

protein expression of active β-catenin over 24 and 48 hours post treatment with phenoxodiol

LNCaP cells exhibited a significant decrease in β-catenin protein versus DMSO vehicle control

after 24 hours of 10µM phenoxodiol treatment (p=0.045) and after 48 hours of treatment with

both 10µM (p=0.023) and 30µM (p=0.046) phenoxodiol. DU145 cells exhibited an increase in

β-catenin expression with the 24 hour 10µM phenoxodiol treatment group increased over

both DMSO vehicle control (p=0.019) and 30µM phenoxodiol (p=0.011), while the 48 hour

30µM phenoxodiol treatment was increased versus DMSO vehicle control (p=0.012). PC3 cells

were similar to the LNCaP cell response with a decrease over 24 hours versus DMSO vehicle

control in both 10µM (p=0.022) and 30µM (p<0.05) phenoxodiol treatment groups.

Page 214: Investigating the effects of, the isoflavone, Phenoxodiol

194

FIGURE 49 ACTIVE Β-CATENIN PROTEIN LEVEL ANALYSIS OF PROSTATE CANCER CELLS OVER 24 AND 48

HOURS POST PHENOXODIOL TREATMENT.

* indicates significance relative to DMSO vehicle control for that time point. ** indicates

significance relative to 30µM Phenoxodiol treatment and DMSO vehicle control for that time

point.

0

0.2

0.4

0.6

0.8

1

1.2

24 Hours 48 Hours

β-c

ate

nin

/β-a

ctin

Treatment

LNCaP active β-catenin Protein Levels

Control

10µM PXD

30µM PXD

0

0.5

1

1.5

2

2.5

3

3.5

24 Hours 48 Hours

β-c

ate

nin

/β-a

ctin

Treatment

DU145 active β-catenin Protein Levels

Control

10µM PXD

30µM PXD

0

0.2

0.4

0.6

0.8

1

1.2

1.4

24 Hours 48 Hours

β-c

ate

nin

/β-a

ctin

Treatment

PC3 active β-catenin Protein Levels

Control

10µM PXD

30µM PXD

Page 215: Investigating the effects of, the isoflavone, Phenoxodiol

195

sFRP4 is an antagonist to the canonical Wnt/β-catenin signalling pathway that interacts with

androgen receptor and can activate Cyclin-D1, and therefore the cell cycle (Horvath et al.

2007; Drake et al. 2009). Figure 50 demonstrates the protein level of sFRP4 over 24 and 48

hours post phenoxodiol treatment in prostate cancer cells. LNCaP cells exhibited a significant

increase in sFRP4 protein expression after treatment with 10µM phenoxodiol over 24

(p=0.0035) and 48 hours (p<0.05) versus control. While the 30µM treatments were not

statistically significant versus control they trended towards biological significance over 24

(p=0.1) and 48 (p=0.066) hours of treatment. DU145 cells exhibited a significant increase in

sFRP4 levels after 24 hours of treatment with 30µM phenoxodiol versus control (p=0.028) and

versus 10µM phenoxodiol (p=0.0072). While no significant difference was determined over 48

hours versus control, a significant difference was detected between the 10µM and 30µM

phenoxodiol treatments (p=0.038). Finally PC3 cells exhibited a significant increase in sFRP4

levels versus control in the 10µM (p=0.021) and 30µM (p=0.008) over 24 hours of treatment.

No significant difference was detected in the 48 hour treatment group of PC3 cells. sFRP4

levels were found to alter after treatment with phenoxodiol with LNCaP, DU145 and PC3 cells

lines exhibiting increases in protein expression following 24 hours of treatment.

Page 216: Investigating the effects of, the isoflavone, Phenoxodiol

196

FIGURE 50 SFRP4 PROTEIN LEVEL ANALYSIS OF PROSTATE CANCER CELLS OVER 24 AND 48 HOURS POST

PHENOXODIOL TREATMENT

* indicates significance relative to DMSO vehicle control for that time point. ** indicates

significance relative to DMSO vehicle control and 10µM Phenoxodiol treatment for that time

point. Ϯ indicates significance relative to 30µM Phenoxodiol only for that point.

0

1

2

3

4

5

24 Hours 48 Hours

sFR

P4

/β-a

ctin

Treatment

LNCaP sFRP4 Protein Levels

Control

10µM PXD

30µM PXD

0

0.2

0.4

0.6

0.8

24 Hours 48 Hours

sFR

P4

/β-a

ctin

Treatment

DU145 sFRP4 Protein Levels

Control

10µM PXD

30µM PXD

0

0.5

1

1.5

2

24 Hours 48 Hours

sFR

P4

/β-a

ctin

Treatment

PC3 sFRP4 Protein Levels

Control

10µM PXD

30µM PXD

Page 217: Investigating the effects of, the isoflavone, Phenoxodiol

197

6.5. DISCUSSION

Two of the hallmarks of cancer are the ability to be self-sufficient in growth signals and to have

infinite replicative potential, which can be initiated through damaging the cell cycle restriction

points, allowing for progression of cancer from benign to metastatic (Hanahan and Weinberg

2000). Here we report that phenoxodiol induces cell cycle arrest in the G1/S phase of the cell

cycle, with the resultant arrest due to the upregulation of p21WAF1. The cytotoxicity may be due

to downstream signalling of molecules such as Akt and ASK1 (Bott et al. 2005; Xie et al. 2010;

Ahmad et al. 2011). c-Myc is a potent oncogene and expression was found to alter in PC3 cells

in response to phenoxodiol. We also report that the expression of Ki-67 and Cyclin-D1 was

altered after phenoxodiol treatment and that the canonical Wnt pathway protein, β-catenin

levels were decreased in LNCaP and PC3 cells while the Wnt receptor antagonist sFRP4 levels,

were increased in all cell lines.

Upon activation of mitogenic signalling, cells commit to entry into a series of regulated steps

allowing completion of the cell cycle. Cells begin in G1 phase, the time between M and S

phases, and before entry into S phase, where DNA is replicated, must pass through a

restriction point (Pardee 1974) that analyses and attempts to repair DNA damage. After S

phase, cells enter G2 phase (the time between the S and M phases) where cells can repair

errors that occurred during DNA duplication, preventing the propagation of these errors to

daughter cells. Finally, the separation into two daughter cells by chromatid separation occurs

and is called M phase (Senderowicz 2004). The sequence of events in cell cycle progression is

highly orchestrated and depends on the cyclic activation and inactivation of cyclin dependent

kinases (CDK), which govern the progression of the cells from one phase to another. In the

event of tumourigenisis, constitutive mitogenic signalling as well as mutations in tumour

Page 218: Investigating the effects of, the isoflavone, Phenoxodiol

198

suppressor genes and proto-oncogenes leads to cell cycle deregulation and uncontrolled

proliferation (MacLachlan et al. 1995; Roy et al. 2007).

The tumour suppressor p53 is the primary controller of cell cycle activity, which triggers cell

cycle arrest, induces the repair of DNA damage or apoptosis by induction of p21WAF1, p53R2,

Bax and Puma (Devlin et al. 2008). p21WAF1 is a cyclin dependant kinase inhibitor (CdkI) family

member, along with p27 and p57, which interfere with the cyclin dependant kinase-cyclin

complex. The group is regulated both by internal and external signals with the expression of

p21WAF1 under transcriptional control of the p53 tumour suppressor gene (Srivastava et al.

2007). In this study we exhibited the ability of phenoxodiol to induce cell cycle arrest at both

10µM and 30µM concentrations over 24 and 48 hours of treatment, with the resultant cells

exhibiting significantly decreased cell populations of G2 cells and subsequent arrest of the cell

cycle visible in the significant alteration of G1 and S phase cell populations. In all cell lines the

decrease in G2 phase cell population was consistent and resulted in very low cell populations,

while some treatments resulted in high S phase arrest and others in G1 phase arrest, it’s clear

that the method of phenoxodiol induced cell cycle arrest is independent of p53 status, with

LNCaP (p53-wild type), DU145 (p53-mutated) and PC3 (p53-null) cells all representing different

p53 status cell types.

p21WAF1 is a tumour suppressor gene that can induce disruption of the Cyclin-e/Cdk2 complex

and prevent subsequent progression from G1 phase into S phase of the cell cycle (el-Deiry et

al. 1993; Bott et al. 2005). Numerous studies have shown that upregulation of p21WAF1 causes

growth arrest in various cancer models and, though p21WAF1 was initially identified to be

transcriptionally up-regulated by p53 in response to DNA damage, recent studies have shown

that p21WAF1 can also be induced by various transcription factors with subsequent mediation of

Page 219: Investigating the effects of, the isoflavone, Phenoxodiol

199

cell cycle arrest, senescence, and apoptosis in a p53 independent manner (Aguero et al. 2005;

Roy et al. 2008). The ability of isoflavones, and specifically phenoxodiol, to induce cell cycle

arrest has been previously reported with studies indicating that arrest was induced by p21WAF1

stabilisation and expression increase (Aguero et al. 2005; Aguero et al. 2010; Seo et al. 2011).

We investigated the expression of p21WAF1 after the data indicated significant cell cycle arrest

as a response to phenoxodiol treatment. p21WAF1 signalling expression was found to be

significantly increased across all the cell lines in response to treatment, indicating that

activation of p21WAF1 was occurring via a p53 independent manner, with resulting cell cycle

arrest. The induction of cytotoxicity in the cells was independent of caspase activation, as

previously shown, and potentiated by induced mitotic depolarisation. This confirms previous

studies that have indicated that isoflavones induce cell cycle arrest through activation and

stabilisation of p21WAF1 (Aguero et al. 2010; Seo et al. 2011).

The assembly of Cyclin-D1, with its CDK4/6 partners, is a mitogen regulated process occurring

in early G1; with the resultant Cyclin-D1-CDK4/CDK6 complexes promoting G1 progression by

inhibiting the activity of the retinoblastoma protein (Rb), resulting in activation of E2F and

subsequent cyclin/cdk signalling, which enters the cell into the cell cycle (Baldin et al. 1993;

Ladha et al. 1998). Many oncogenic signals induce Cyclin-D1 expression and do so through

distinct DNA sequences in the Cyclin-D1 promoter, including Ras, Src,ErbB2, and β-catenin.

Decreasing the expression of Cyclin-D1, or interference with the cyclin/Cdk complex results in

cell arrest in G1 phase and eventual senescence. We investigated the expression of Cyclin-D1

after treatment with phenoxodiol and determined that DU145 and PC3 both had a significant

decrease in signalling over 24 hours but not over 48, while LNCaP cells did not change

expression of cyclin-D1 signalling. While not a direct target of phenoxodiol treatment in

prostate cancer cells, the role of Cyclin-D1 seems to be tissue and oncogene specific, with

Cyclin-D1 linked to activation of the Wnt/β-catenin signalling pathway (Fu et al. 2004).

Page 220: Investigating the effects of, the isoflavone, Phenoxodiol

200

Ki-67 antigen is present in all proliferating cells (normal and neoplastic) and its evaluation

allows determination of the rate of growth. Ki-67 expression has been shown to have a strong

relationship with Gleason's grading, which has an important correlation with the prognosis of

prostate cancer and, as such, it is an independent predictive factor in both patient survival

assessment and disease recurrence (Koda et al. 2007; Madani et al. 2011). Ki-67 expression

was found to significantly decrease in LNCaP and PC3 cells, while there was a biological trend

towards this in DU145 cells but error margins resulted in no detection of significant alteration.

The decreased Ki-67 signalling expression confirms the cell cycle arrest data and indicates that

prostate cancer cells are undergoing senescence induced cytotoxicity. c-Myc signalling has

been shown to be a proto-oncogene, with regulation of c-Myc expression inducing the

expression of other oncogenes, in response and leading to a neoplastic cell type. It is known

that cell cycle regulation is altered under excess c-Myc expression, with a decrease in time

taken to reach the restriction point of G1 (Wasylishen and Penn 2010). In this study we

determined that c-Myc expression was only altered under the influence of phenoxodiol in PC3

cells after 48 hours of treatment, with signalling potentially being a mechanism to drive the

cell out of arrest.

Another controller of the cell cycle is the canonical Wnt/β-catenin pathway where Wnt growth

signals cause decreased expression of GSK3β protein, which results in an active form of β-

catenin translocating to the nucleus and initiating cell cycle progression through multiple

pathways including Cyclin-D1 and c-Myc expression (Zhang et al. 2011). An up regulated

canonical Wnt/β-catenin pathway has multiple functions; it can be potentiated by AR

expression and has been implicated in both cell adhesion and signal transduction in response

to Wnt ligands, resulting in an invasive metastatic phenotype (Chesire et al. 2002). We

investigated the expression of the active form of β-catenin and determined that LNCaP and

PC3 cells had significantly reduced active β-catenin expression after 24 hours of phenoxodiol

Page 221: Investigating the effects of, the isoflavone, Phenoxodiol

201

treatment, with LNCaP continuing on to a significant 48 hour effect, which coincides with a

subsequent decrease in Cyclin-D1 signalling expression in androgen insensitive PC3 cells while

LNCaP expression was unchanged due to the potential interaction of AR receptor. DU145 cells

exhibited an increased expression of active β-catenin after treatment over 24 and 48 hours,

indicating that alteration of canonical Wnt signalling is not a direct effect of phenoxodiol

treatment but a downstream signalling outcome.

Secreted frizzled related protein 4 (sFRP4) is one of a group of proteins that antagonise the

canonical Wnt/β-catenin pathway. The interference with Wnt activity results in decreased

activated β-catenin formation and decreased invasive potential. sFRP4 has been verified to

have an anti-angiogenic ability as well as the ability to induce apoptotic signalling (Drake et al.

2009; Muley et al. 2010). sFRP4 appears to affect androgen-dependent and androgen-

independent prostate cancer and has a role as a prognostic marker for androgen dependant

prostate cancer (Horvath et al. 2007). We determined that sFRP4 expression was increased

over 24 hours in PC3 cells, coinciding with decreased active β-catenin levels in that cell line and

increased levels over 24 and 48 hours in LNCaP cells, coinciding with decreased active β-

catenin in that cell line. sFRP4 levels were increased in DU145 cells but did not correspond

with β-catenin changes, indicating that DU145 cells might have an alteration in the canonical

Wnt/β-catenin signalling pathway. Phenoxodiol exhibited the potential to regulate canonical

Wnt/β-catenin signalling through increased expression of sFRP4 and subsequent decreased

expression of active β-catenin in two of the cell lines. The functionality of this canonical Wnt/β-

catenin antagonism in prostate cancer must be determined before treatment with

phenoxodiol can be utilised for successful targeting of the pathway directly. However, soy

isoflavone interaction with the pathway has previously been exhibited (Liss et al. 2010).

Page 222: Investigating the effects of, the isoflavone, Phenoxodiol

202

In this study we determined that phenoxodiol treatment induced significant cell cycle arrest

across 24 and 48 hours of 10µM and 30µM phenoxodiol treatment. p21WAF1 signalling

expression was found to be significantly increased across all the cell lines in response to

treatment, indicating that activation of p21WAF1 and subsequent cell arrest was occurring via a

p53 independent manner, with induction of cytotoxicity independent of caspase activation.

We determined that c-Myc and Cyclin-D1 expression was not consistently altered but that Ki-

67 signalling expression was decreased in line with the cell cycle arrest. The interaction of

sFRP4 and active β-catenin implied that phenoxodiol treatment results in an alteration of

sFRP4 and β-catenin signalling though changes are dependent on the cell line. Phenoxodiol

demonstrates an ability in prostate cancer cells to induce significant cytotoxicity in cells by

interacting with p21WAF1 and inducing cell cycle arrest irrespective of p53 status or caspase

pathway interactions. These data indicate that phenoxodiol would be effective as a potential

future treatment modality for both hormone sensitive and hormone refractory prostate

cancer.

Page 223: Investigating the effects of, the isoflavone, Phenoxodiol

203

7 . C H A P TER S EV EN : P H E NOX OD IOL IN C OMB INA TION W ITH D OC ETA X EL

7.1. INTRODUCTION

Advanced prostate cancer has a 5-year survival of only 30%. Historically, chemotherapy has

been used with palliative intent but unclear survival benefit for these advanced-stage patients

(Montero et al. 2005). The lack of curative measures for late stage prostate cancer and the

ability to develop resistance to treatment have been linked to the existence of cancer stem

cells. The cancer stem cell population acts as an initiator of cancer and exists in low numbers.

Cancer stem cells have the ability to acquire resistance upon exposure to treatment, which

results in the development of a new population of cells that are now resistant to previously

successful treatment (Reya et al. 2001; Trosko et al. 2004). Current practices for hormone-

refractory/castrate resistant, metastatic prostate cancer incorporate the use of taxanes.

Docetaxel, in particular, is being incorporated in numerous current clinical trials either as a

single or combination agent against androgen-independent prostate cancer, and it is also

being investigated for its use as a neoadjuvant or adjuvant agent in hormone sensitive, locally

aggressive prostate cancer (Canfield et al. 2006).

Docetaxel and paclitaxel are from a class of anti-cancer agents called taxanes that bind to and

stabilise microtubules causing G2/M cell-cycle arrest and apoptosis. Taxanes have a different

mechanism of action from that of any other class of anti-cancer drugs, namely hyper-

stabilisation of microtubules (Montero et al. 2005). Taxanes function by targeting the subunit

of the tubulin heterodimer, the key component of cellular microtubules that allow for

chromatid separation during the M cell cycle phase. Although the action and anti-cancer

activity of paclitaxel and docetaxel are much the same, key differences exist clinically;

Page 224: Investigating the effects of, the isoflavone, Phenoxodiol

204

docetaxel shows activity in patients with metastatic solid tumours that are resistant to

paclitaxel (Michaud et al. 2000). The actual mechanisms that lead to cell death remain unclear

but may include activation of intrinsic pathways essential for apoptosis, induction of bcl-2

phosphorylation facilitating apoptosis and inhibition of angiogenesis. Cell death after exposure

to docetaxel appears to involve apoptotic mechanisms, including classic features such as DNA

fragmentation, cell volume shrinkage, and membrane-bound apoptotic bodies (Michaud et al.

2000). Studies have also shown that apoptosis induced by taxanes involves several apoptotic

signal molecules, such as JNK, protein kinase A, c-Raf-1/Ras/Bcl-2, p53/p21WAF1 and mitogen-

activated protein kinases (ERK and p38). However, the mode of apoptotic action in different

tumours is far from clear (Gan et al. 2011).

Combination therapies have the potential to increase the effectiveness of drug treatments

while simultaneously increasing quality of life by reducing side effects, lowering effective

dosage rates or by increasing effectiveness of one compound once combined with another.

Successful combination therapies that produce an increase in quality of life for patients, even if

not associated with life extension, are a highly valued outcome of research, even with drugs

that do not directly interact (Kantoff et al. 1999; Parente et al. 2012). An example of this is the

use of Docetaxel with androgen ablation therapy in late stage prostate cancer defined as

castrate resistant prostate cancer (CRPC). To date, docetaxel-based chemotherapy remains the

only treatment that has demonstrated an overall survival benefit in most men with metastatic

CRPC regardless of whether they are symptomatic or have visceral metastases (Hotte and Saad

2010). Despite the initial effectiveness of these treatments, late stage prostate cancer

treatment is still palliative and characterised by androgen independence, metastasis and

poorly differentiated cells.

Page 225: Investigating the effects of, the isoflavone, Phenoxodiol

205

To determine an effective combination therapy accurately, the interaction of the two

compounds is analysed and the effectiveness investigated for synergistic, additive or even

interference effects (Zhao et al. 2004). True synergism is rare as most drugs in combination are

additive. Due to clearance rates of these drugs through the body keeping serum levels at an

effective concentration means that, in an additive environment, one might not be able to

lower effective concentrations, thereby increasing the risk of side effects but theoretically

decreasing the length of time treatment must be carried out (Tallarida 2006). Isobolograms are

one method of determining effective ranges of drug treatments in combination, because

multiple concentrations of drugs can be compared against each other and a profile of effective

dose rates can be constructed for the environment being tested. In prostate cancer surgical

intervention, radiotherapy and chemotherapy are utilised, sometimes in combination, to treat

tumours.

Cytoplasmic and/or nuclear β-catenin can be used as an indicator of activation of the Wnt/β-

catenin pathway because it is observed in up to 71% of advanced prostate tumour specimens

(Chesire et al. 2002). Increasing evidence also suggests that canonical Wnt signalling could

function to assist in bone metastasis formation, a progressive state that has a particularly poor

prognosis (Hall et al. 2006). It has been revealed that the Wnt ligand target, the frizzled

receptor, is over expressed in prostate cancer but that this over expression is counterbalanced

by the secreted frizzled related protein (sFRP) family which attempts to suppress AR-mediated

transactivation (Kawano et al. 2009).

Methylation of the sFRP proteins has been suggested as a marker of invasive carcinoma, with a

resulting poor prognosis (Ahmad et al. 2011). Apart from β-catenin degradation, it has been

suggested that sFRP4 inhibition of Wnt signalling causes stabilisation of GSK3β, resulting in p53

Page 226: Investigating the effects of, the isoflavone, Phenoxodiol

206

activation and associated signalling such as p21WAF1, as well as a decreased invasive potential in

androgen-independent prostate cancer cells. sFRP4 has also been indicated to have

antiangiogenic and pro-apoptotic properties in a wide variety of cells (Watcharasit et al. 2002;

Fox and Dharmarajan 2006; Muley et al. 2010). Unlike other inhibitors of Wnt signalling, sFRP4

appears to affect androgen-dependent and androgen-independent prostate cancer (Horvath et

al. 2007).

In this study we seek to investigate the cytotoxic effects of the prostate cancer treatment

docetaxel, and determine a range of treatment concentrations. Potential interactions of

phenoxodiol and docetaxel across multiple concentrations are investigated, searching for

synergistic, additive or interference effects. We also investigate if any cytotoxic effects are

increased through pretreatment with docetaxel or phenoxodiol and, finally, we try to

determine if the Wnt/β-catenin antagonist sFRP4, when combined with phenoxodiol, will have

an interactive effect.

Page 227: Investigating the effects of, the isoflavone, Phenoxodiol

207

7.2. AIMS

The aims of this chapter were as follows:

Aim 1: To characterise the cytotoxic effect of Docetaxel on prostate cancer cells.

Aim 2: To determine the effect of combination therapy with docetaxel and phenoxodiol on

prostate cancer cells.

Aim 3: To determine if pretreatment with phenoxodiol or docetaxel has an impact on the

effect of subsequent combination therapy.

Aim 4: To determine the effect of combination therapy with phenoxodiol and sFRP4 purified

protein on prostate cancer cells.

7.3. METHODOLOGY

The methodology utilised in this chapter is discussed in detail in the materials and methods

chapter (page 50). Initially, a dose response curve to docetaxel would be needed to determine

effective dose rates to use in conjunction with phenoxodiol. Briefly, cells were seeded onto 96

well plates at appropriate rates and incubated for 48 hours. Media were then aspirated and

replaced with a 10 fold serially diluted concentration of docetaxel from 10µM to .1nM, with

DMSO as the vehicle control, and all samples having an equal concentration of DMSO vehicle

added. After 48 hours of treatment, 20µL of MTS dye were added and the cells incubated for 3

hours before absorbance was measured at 492nm. Increased absorbance was linked to

increased cell metabolism and, therefore, proliferation. No interaction was noted between

docetaxel and the MTS dye.

Page 228: Investigating the effects of, the isoflavone, Phenoxodiol

208

An isobologram was performed between docetaxel and phenoxodiol; the samples were all

balanced for vehicle concentration. Briefly, cells were seeded into 96 well plates and incubated

for 48 hours; afterwards they had the cell media aspirated and a treatment media applied,

containing any combination listed in Table 23, resulting in 24 different treatments.

TABLE 23 PHENOXODIOL/DOCETAXEL CONCENTRATION COMBINATIONS FOR AN ISOBOLOGRAM

Drug Name Conc.

Phenoxodiol 0µM 5µM 10µM 30µM

Docetaxel 0nM .1nM 1nM 5nM 10Nm 100nM

A list of potential concentration combinations of Phenoxodiol and Docetaxel during the

isobologram, resulting in a final quantity of 24 individual treatment types, all balanced for

vehicle control.

The isobologram was performed after 48 hours of incubation with treatment media; 20µL of

MTS dye were then added to each well and incubated for 3 hours. Afterwards media

absorbance was measured at 492nm, with the resulting absorbance indicating cell

proliferation rates within the prostate cancer cell lines.

A pretreatment study was then performed. Briefly, cells were seeded at appropriate rates into

96 well plates and incubated for 48 hours, treatment media were applied to each well and

incubation occurred for another 24 hours. After 24 hours, a set of 96 well plates had 20µL of

MTS dye added, were incubated for 3 hours and absorbance was measured at 492nm on a

plate reader. The other plates had the addition of new treatment compounds added to

appropriate wells, i.e. some cells had exposure to 24 hours of 30µM phenoxodiol then 24

Page 229: Investigating the effects of, the isoflavone, Phenoxodiol

209

hours of exposure to 100nM docetaxel and 30µM phenoxodiol in combination. This tested for

potential induction of pre-treatment sensitivity. After 48 hours of total treatment with the

initial compound concentration, the remaining 96 well plates had 20µL of MTS dye added and

were incubated for 3 hours, followed by measurement of absorbance at 492nm on a plate

reader.

Finally, a Wnt antagonist and phenoxodiol treatment study was performed. Briefly, cells were

seeded into 96 well plates at appropriate rates and incubated for 48 hours, after which cells

were treated with varying concentrations of purified sFRP4 protein up to 500pg/mL, which was

combined with 30µM phenoxodiol treatment. Vehicle consisted of PBS and DMSO

respectively. Following 48 hours of exposure to treatment, cells had 20µL of MTS dye added

and were incubated for 3 hours before absorbance was measured at 492nm on a plate reader.

Page 230: Investigating the effects of, the isoflavone, Phenoxodiol

210

7.4. RESULTS

Docetaxel is a common treatment for early and late stage prostate cancer, and is highly

cytotoxic as a cell microtubulin stabiliser, causing G2 phase cell cycle blockage in replicating

cells by damaging the mitotic spindle (Michaud et al. 2000). As phenoxodiol is a novel drug for

prostate cancer with an ability to affect the cell cycle, a series of experiments were run to

determine potential synergy, additivity, interference or lack of interaction between each

compound.

Docetaxel is known to induce G2 phase cell cycle arrest by inhibiting the activity of the mitotic

spindle (Montero et al. 2005). Figure 51 demonstrates the cell proliferation rates of prostate

cancer cell lines after 48 hours of treatment with 10 fold dilutions of docetaxel spanning .1nM

to 10µM concentrations. LNCaP cells exhibited a significant increase in cell proliferation versus

DMSO vehicle control after treatment with 1nM docetaxel (p≤0.001). LNCaP cells treated with

10nM, 100nM, 1µM, and 10µM concentrations all exhibited a significant decrease in

proliferation versus control (p=0.033, p≤0.001 for the last 3). While 100nM, 1µM, and 10µM

concentrations were not significantly different to each other they were all significantly

decreased over the 10nM docetaxel concentration (p≤0.001 for all).

DU145 cells had a similar pattern of response to phenoxodiol treatment with .1nM and 1nM

doses significantly increasing cell proliferation versus control (p≤0.001 for all). While 10nM,

100nM, 1µM, and 10µM concentrations exhibited a significant decrease in proliferation versus

control (p≤0.001 for all). There was no significant difference detected between 100nM, 1µM,

and 10µM docetaxel treatments but all three were significantly decreased with respect to the

10nM treatment (p≤0.001 for all).

Page 231: Investigating the effects of, the isoflavone, Phenoxodiol

211

PC3 cells exhibited a similar trend with .1nM and 1nM docetaxel treatments significantly

increasing cell proliferation versus control (p≤0.001 for both), and 10nM, 100nM, 1µM and

10µM docetaxel concentrations significantly decreasing proliferation versus control (p≤0.001

for all). Only the 1µM docetaxel concentration was found to be significantly decreased versus

10nM and 100nM concentrations (p=0.042). All three cell lines exhibited a significant decrease

in proliferation in response to high concentrations of docetaxel, as such the 100nM dose of

docetaxel was chosen as the lowest effective dose over 48 hours of treatment.

Page 232: Investigating the effects of, the isoflavone, Phenoxodiol

212

FIGURE 51 DOCETAXEL RESPONSE CURVE MEASURED AFTER 48 HOURS OF TREATMENT.

* indicates significance relative to control after 48 hours. ** indicates significance relative to

control and * groups after 48 hours. *** indicates significance relative to control and *, **

groups after 48 hours.

0

20

40

60

80

100

120

140

% C

ell

Gro

wth

ve

rsu

s C

on

tro

l

Treatment

LNCaP Docetaxel Response over 48 Hours

Control .1nM 1nM 10nM 100nM 1µM 10µM

0

20

40

60

80

100

120

140

% C

ell

Gro

wth

ve

rsu

s C

on

tro

l

Treatment

DU145 Docetaxel Response over 48 Hours

Control .1nM 1nM 10nM 100nM 1µM 10µM

0

20

40

60

80

100

120

140

% C

ell

Gro

wth

ve

rsu

s C

on

tro

l

Treatment

PC3 Docetaxel Response over 48 Hours

Control .1nM 1nM 10nM 100nM 1µM 10µM

Page 233: Investigating the effects of, the isoflavone, Phenoxodiol

213

Phenoxodiol has exhibited the ability to induce G1/S phase cell cycle arrest, while docetaxels

published method of action is to inhibit the cell cycle at the M phase, resulting in G2 phase

arrest (Michaud et al. 2000; Aguero et al. 2010). After determining the minimum effective

concentration dose an isobologram was performed to look for interaction between

phenoxodiol and docetaxel with varying concentrations of each utilised. Figure 52

demonstrates an isobologram analysis of phenoxodiol and docetaxel treatments, over 48

hours, with a total of 24 individual treatments all standardised to DMSO vehicle control in

LNCaP cells. Isobolograms determine interference, synergy and additivity with significant

interference effects indicated in the follow graphs. Additive effects were evident throughout

the isobologram but, while there was a significant decrease in cell proliferation rate versus

control (0µM PXD, 0nM DOC) over most populations, there was a significant increase in

proliferative rate versus 100nM docetaxel only for that time point (p=0.014) indicating that the

5µM phenoxodiol treatment inhibited the effectiveness of 100nM docetaxel concentration.

These data provide evidence for phenoxodiol interfering with LNCaP cell docetaxel treatment

at a specific concentration combination.

FIGURE 52 LNCAP PHENOXODIOL / DOCETAXEL ISOBOLOGRAM MEASURED AFTER 48 HOURS OF

TREATMENT

0

20

40

60

80

100

120

140

0nM DOC .1nM DOC 1nM DOC 5nM DOC 10nM DOC 100nM DOC

% C

ell

Gro

wth

vs

Co

ntr

ol

LNCaP Phenoxodiol/Doxetaxel Isobologram

0µM PXD 5µM PXD 10µM PXD 30µM PXD

Page 234: Investigating the effects of, the isoflavone, Phenoxodiol

214

* indicates significant increase in proliferation relative to [0µM PXD, 100nM DOC] treatment

after 48 hours.

Figure 53 demonstrates an isobologram analysis of phenoxodiol and docetaxel treatments

over 48 hours with a total of 24 individual treatments all standardised to DMSO vehicle control

in DU145 cells. Additive effects were evident throughout the isobologram but, while there was

a significant decrease in cell proliferation rate versus control (0µM PXD, 0nM DOC) over most

populations, the indicated samples displayed a significant increase in proliferative rate versus

100nM docetaxel only for that time point (5µM PXD p≤0.001, 10µM PXD p=0.015), indicating

that the 5µM and 10µM concentrations of phenoxodiol inhibited the effectiveness of a 100nM

docetaxel concentration. These data provide evidence for phenoxodiol interfering with DU145

cell docetaxel treatment at a specific concentration combination.

FIGURE 53 DU145 PHENOXODIOL / DOCETAXEL ISOBOLOGRAM MEASURED AFTER 48 HOURS OF

TREATMENT.

* indicates significant increase in proliferation relative to [0µM PXD, 100nM DOC] treatment

after 48 hours.

0

20

40

60

80

100

120

140

0nM DOC .1nM DOC 1nM DOC 5nM DOC 10nM DOC 100nMDOC

% C

ell

Gro

wth

vs

Co

ntr

ol

DU145 Phenoxodiol/Docetaxel Isobologram

0µM PXD 5µM PXD 10µM PXD 30µM PXD

Page 235: Investigating the effects of, the isoflavone, Phenoxodiol

215

Figure 54 demonstrates an isobologram analysis of phenoxodiol and docetaxel treatments

over 48 hours with a total of 24 individual treatments all standardised to DMSO vehicle control

in PC3 cells. Additive effects were evident throughout the isobologram but, while there was a

significant decrease in cell proliferation rate versus control (0µM PXD, 0nM DOC) over most

populations, the indicated samples displayed a significant increase in proliferative rate versus

100nM docetaxel only for that time point (p=0.015), indicating that the 5µM phenoxodiol

treatment inhibited the effectiveness of 100nM docetaxel only concentration. These data

provide evidence for phenoxodiol interfering with PC3 cell docetaxel treatment at a specific

concentration combination.

FIGURE 54 PC3 PHENOXODIOL / DOCETAXEL ISOBOLOGRAM MEASURED AFTER 48 HOURS OF

TREATMENT.

* indicates significant increase in proliferation relative to [0µM PXD, 100nM DOC] treatment

after 48 hours.

Following the isobologram results, a more in depth investigation of potential sensitisation

needed to be performed with pre-treatment of the cells with phenoxodiol or docetaxel and

0

20

40

60

80

100

120

140

0nM DOC .1nM DOC 1nM DOC 5nM DOC 10nM DOC 100nMDOC

% C

ell

Gro

wth

vs

Co

ntr

ol

PC3 Phenoxodiol/Docetaxel Isobologram

0µM PXD 5µM PXD 10µM PXD 30µM PXD

Page 236: Investigating the effects of, the isoflavone, Phenoxodiol

216

then subsequent addition of the remaining agent. Figure 55 demonstrates a combination

therapy treatment of 10µM phenoxodiol and 100nM docetaxel in prostate cancer cells over 24

and 48 hours, with two samples receiving 24 hours solo drug treatment and then the next 24

hours of combined treatment. LNCaP cells exhibited a significant decrease in proliferation

versus control over 24 hours in the 10µM PXD (p=0.0089) and combination (p<0.001)

treatments, while combination therapy was also significantly lower than the 10µM PXD

treatment (p=0.0096). LNCaP cells exhibited a significant decrease in proliferation versus

control over 48 hours when 10µM PXD (p<0.001), combination (p<0.001) and 48Hr PXD

(p<0.001) treatments were used. The 100nM DOC (p<0.05) and 48Hr DOC (p=0.0035) were all

significantly decreased in proliferation versus 48Hr PXD as well as 10µM PXD combination and

Control treatments. Lastly, 48hr DOC was also significantly decreased versus 100nM DOC

(p=0.022).

DU145 cells exhibited a significant decrease in proliferation over 24 hours with 10µM PXD

(p<0.001), 100nM DOC (p<0.001) and combination (p<0.001) treatments. After 48 hours,

DU145 cells exhibited significantly decreased proliferation versus control in 10µM PXD

(p<0.001) and 48Hr PXD (p<0.001) treatments but exhibited no significant difference between

each other, while 100nM DOC (p=0.022), combination (p=0.027) and 48Hr DOC (p=0.019)

treatments were all significantly decreased versus 48Hr PXD as well as 10µM PXD and Control.

PC3 cells exhibited a significant decrease in proliferation versus control after 24 hours of

treatment with 10µM PXD (p<0.001), 100nM DOC (p<0.001) and combination (p<0.001)

treatments, while 100nM DOC (p=0.0072) and combination (p=0.024) treatments were also

significantly decreased versus 10µM PXD. Following the trend set by LNCaP and DU145 cells,

PC3 cells exhibited a significant decrease in proliferation versus control over 48 hours after

Page 237: Investigating the effects of, the isoflavone, Phenoxodiol

217

treatment with 48Hr PXD (p<0.0011), while 10µM PXD (p=0.0062), 100nM DOC (p=0.0041),

combination (p=0.0026) and 48Hr DOC (p=0.0029) exhibited significantly decreased

proliferation versus 48Hr PXD as well as Control.

* indicates significance relative to control for that time point. ** indicates significance relative

to control and * groups for that time point. *** indicates significance relative to control and *,

** groups for that time point.

The [48Hr 10µM PXD+24Hr 100nM DOC] and [48Hr 100nM DOC+24Hr 10µM PXD] treatments

will be referred to as the 48hr PXD and 48Hr DOC treatments respectively while the [48Hr

10µM PXD+48Hr 100nM DOC] will be referred to as the combination treatment.

Page 238: Investigating the effects of, the isoflavone, Phenoxodiol

218

FIGURE 55 10µM PHENOXODIOL 100NM DOCETAXEL COMBINATION THERAPY MEASURED AFTER 48

HOURS OF TREATMENT.

0

20

40

60

80

100

24 Hours 48 Hours

% C

ell

Gro

wth

vs

Co

ntr

ol

Treatment

LNCaP 10µM PXD/DOC Combined Treatment

Control

10µM PXD

100nM DOC

10µM PXD + 100nM DOC

48Hr 10µM PXD+24Hr100nM DOC

48Hr 100nM DOC+24Hr10µM PXD

0

20

40

60

80

100

24 Hours 48 Hours

% C

ell

Gro

wth

vs

Co

ntr

ol

Treatment

DU145 10µM PXD/DOC Combined Treatment

Control

10µM PXD

100nM DOC

10µM PXD + 100nM DOC

48Hr 10µM PXD+24Hr100nM DOC

48Hr 100nM DOC+24Hr10µM PXD

0

20

40

60

80

100

24 Hours 48 Hours

% C

ell

Gro

wth

vs

Co

ntr

ol

Treatment

PC3 10µM PXD/DOC Combined Treatment

Control

10µM PXD

100nM DOC

10µM PXD + 100nM DOC

48Hr 10µM PXD+24Hr100nM DOC

48Hr 100nM DOC+24Hr10µM PXD

Page 239: Investigating the effects of, the isoflavone, Phenoxodiol

219

Figure 56 Indicates a combination therapy treatment of 30µM phenoxodiol and 100nM

docetaxel in prostate cancer cells, over 24 and 48 hours, with two samples receiving 24 hours

solo drug treatment then the next 24 hours of combined treatment. Phenoxodiol has exhibited

the ability to induce G1/S phase cell cycle arrest, and docetaxel’s published method of action is

to inhibit the cell cycle at the M phase, resulting in G2 phase arrest (Michaud et al. 2000;

Aguero et al. 2010). LNCaP cells did not exhibit a significant change in proliferation versus

control over 24 hours. LNCaP cells exhibited a significant decrease in proliferation versus

control over 48 hours, when 30µM PXD and 48Hr PXD treatments were used (p<0.05,

p<0.033). The 100nM DOC (p=0.012), combination (p=0.0092) and 48Hr DOC (p=0.0090) were

all significantly decreased in proliferation versus 48Hr PXD, as well as 30µM PXD and Control

treatments..

DU145 cells exhibited a significant decrease in cell proliferation versus control in 30µM PXD

(p<0.0034), 100nM DOC (p<0.001) and combination (p<0.001), with both 100nM DOC

(p<0.001) and combination (p<0.005) therapy exhibiting further decreased proliferation versus

30µM PXD. A similar trend occurred in DU145 cells as in LNCaP, with 30µM PXD (p<0.001) and

48Hr PXD (p<0.001) significantly decreased in proliferation versus control but exhibiting no

significant difference between each other, while 100nM DOC, combination and 48Hr DOC

treatments were all significantly decreased versus 48Hr PXD (P<0.001 for all) as well as 30µM

PXD and Control.

Finally, PC3 cells exhibited a significant decrease in cell proliferation versus DMSO vehicle

control over 24 hours for 100nM DOC (p<0.001) and combination therapy (p<0.017), with

100nM DOC also being significantly decreased versus 30µM PXD (p<0.0021) and combined

treatment (p<0.015). Over 48 hours the trend displayed by LNCaP and DU145 continued, PC3

Page 240: Investigating the effects of, the isoflavone, Phenoxodiol

220

cells were significantly decreased in proliferation over time versus control in the 30µM PXD

(p<0.001) and 48Hr PXD (p=0.001) treatments, while 100nM DOC (p=0.0015), combination

(p=0.011) and 48Hr DOC (p=0.0015) were all significantly decreased versus 48Hr PXD as well as

30µM PXD and Control. This data suggests that, in all three cell lines, pre-treatment with

phenoxodiol for 24 hours before the addition of docetaxel to make a combination with

phenoxodiol does not result in a decrease in cell proliferation versus treatment with 30µM

phenoxodiol by itself. These data also suggest that combined treatment with phenoxodiol does

not increase the activity of docetaxel treatment in comparison to 100nM only docetaxel

treatment over 48 hours.

* indicates significance relative to control for that time point. ** indicates significance relative

to control and * groups for that time point.

The [48Hr 30µM PXD+24Hr 100nM DOC] and [48Hr 100nM DOC+24Hr 30µM PXD] treatments

will be referred to as the 48hr PXD and 48Hr DOC treatments respectively while the [48Hr

30µM PXD+48Hr 100nM DOC] will be referred to as the combination treatment.

Page 241: Investigating the effects of, the isoflavone, Phenoxodiol

221

FIGURE 56 30µM PHENOXODIOL 100NM DOCETAXEL COMBINATION THERAPY MEASURED AFTER 48

HOURS OF TREATMENT.

0

20

40

60

80

100

24 Hours 48 Hours

% C

ell

Gro

wth

vs

Co

ntr

ol

Treatment

LNCaP 30µM PXD/DOC Combined Treatment

Control

30µM PXD

100nM DOC

30µM PXD + 100nM DOC

48Hr 30µM PXD+24Hr100nM DOC

48Hr 100nM DOC+24Hr30µM PXD

0

20

40

60

80

100

24 Hours 48 Hours

% C

ell

Gro

wth

vs

Co

ntr

ol

Treatment

DU145 30µM PXD/DOC Combined Treatment

Control

30µM PXD

100nM DOC

30µM PXD + 100nM DOC

48Hr 30µM PXD+24Hr100nM DOC

48Hr 100nM DOC+24Hr30µM PXD

0

20

40

60

80

100

24 Hours 48 Hours

% C

ell

Gro

wth

vs

Co

ntr

ol

Treatment

PC3 30µM PXD/DOC Combined Treatment

Control

30µM PXD

100nM DOC

30µM PXD + 100nM DOC

48Hr 30µM PXD+24Hr100nM DOC

48Hr 100nM DOC+24Hr30µM PXD

Page 242: Investigating the effects of, the isoflavone, Phenoxodiol

222

After determining the effects of pre-treatment with another known cell cycle inhibitor,

Docetaxel, it was decided to test the effect of decreasing Wnt regulatory pathway signalling by

addition of sFRP4 whole protein in cell culture in conjunction with phenoxodiol treatment. This

was to determine the effects of sFRP4 on prostate cancer cells lines and detect any; synergy,

additivity, interference or lack of effect when sFRP4 was used in conjunction with phenoxodiol.

Figure 57 Indicates the effect of the combination of 500pg/mL of secreted frizzled related

protein 4 (sFRP4) and 30µM phenoxodiol treatment, over 48 hours, on prostate cancer cell

lines. Due to the large amount of statistics generated comparing the samples, a general

(p<0.05) will be used to indicate significant differences and only specific differences noted.

LNCaP cells treated with 125pg/mL and 250pg/mL sFRP4 were found to exhibit significantly

reduced proliferation versus control (p<0.05) over 48 hours. LNCaP cells treated with

500pg/mL sFRP4 and 10µM PXD treatments were not determined to be significantly different

to each other but were both significantly reduced in comparison to control, 125pg/mL and

250pg/mL sFRP4 treatments (p<0.05). LNCaP cells treated with 30µM phenoxodiol exhibited

significantly reduced proliferation against control, all sFRP4 concentrations and 10µM

phenoxodiol treatment (p<0.05); and, finally, the combination of 30µM PXD and 500pg/mL

sFRP4 proved to significantly reduce proliferation in comparison to control and all other

treatments (p<0.05), with a large improvement over 30µM phenoxodiol only.

DU145 cells exhibited a significant decrease in proliferation versus control with the 250pg/mL

sFRP4 treatment group (p<0.05), while neither 125pg/mL nor 500pg/mL sFRP4 were

determined to exhibit any significant changes versus control over 48 hours. DU145 cells

treated with either 10µM and 30µM phenoxodiol treatments were significantly decreased in

proliferation in comparison to control and all sFRP4 concentrations (p<0.05) but not versus

Page 243: Investigating the effects of, the isoflavone, Phenoxodiol

223

each other. Finally, the DU145 cells treated with combined sFRP4 and phenoxodiol exhibited a

significant decrease in proliferation versus control and the other treatment groups (p<0.05).

PC3 cells exhibited a significant decrease in cell proliferation versus control, over 48 hours,

with the 10µM phenoxodiol treatment (p<0.05). None of the sFRP4 treatments were

significantly different versus control in the chemoresistant late stage PC3 prostate cancer cells.

PC3 cells treated with 30µM phenoxodiol exhibited significantly reduced proliferation versus

control, all sFRP4 treatments and 10µM phenoxodiol (p<0.05); and finally, similar to the

previous two cell lines, PC3 cells treated with 30µM phenoxodiol and 500pg/mL sFRP4 were

significantly reduced in proliferation in comparison to control and all other treatments. This

study indicates that the combination of 30µM phenoxodiol and 500pg/mL sFRP4 results in a

significant decrease in cell proliferation across all three cell lines. This indicates that sFRP4 may

act as a sensitisation agent to the demonstrated cytotoxic effect of phenoxodiol.

* indicates significance relative to DMSO/PBS vehicle control for that time point.

** indicates significance relative to control and * groups for that time point.

*** indicates significance relative to control, * and ** groups

ϯ is significance relative to all other treatment groups.

Page 244: Investigating the effects of, the isoflavone, Phenoxodiol

224

FIGURE 57 30µM PXD AND 500PG/ML SFRP4 PROTEIN COMBINATION THERAPY AFTER 48 HOURS.

0

20

40

60

80

100

120

% C

ell

Gro

wth

ve

rsu

s C

on

tro

l

Treatment

LNCaP PXD/sFRP4 Combined Treatment Control

125pg/ml sFRP4

250pg/ml sFRP4

500pg/ml sFRP4

10uM PXD

30uM PXD

30uM PXD +500pg/ml sFRP4

0

20

40

60

80

100

120

% C

ell

Gro

wth

ve

rsu

s C

on

tro

l

Treatment

DU145 PXD/sFRP4 Combined Treatment Control

125pg/ml sFRP4

250pg/ml sFRP4

500pg/ml sFRP4

10uM PXD

30uM PXD

30uM PXD +500pg/ml sFRP4

0

20

40

60

80

100

120

140

% C

ell

Gro

wth

ve

rsu

s C

on

tro

l

Treatment

PC3 PXD/sFRP4 Combined Treatment Control

125pg/ml sFRP4

250pg/ml sFRP4

500pg/ml sFRP4

10uM PXD

30uM PXD

30uM PXD +500pg/ml sFRP4

Page 245: Investigating the effects of, the isoflavone, Phenoxodiol

225

7.5. DISCUSSION

Castrate resistant prostate cancer is defined by disease progression despite androgen-

deprivation therapy and may present as one or any combination of a continuous rise in serum

levels of prostate-specific antigen (PSA), progression of pre-existing disease or the appearance

of new metastases. Current practices for hormone-refractory/castrate resistant, metastatic

prostate cancer incorporate the use of taxanes (Petrylak et al. 2004; Tannock et al. 2004). In

this study we investigated the effects of docetaxel on the prostate cancer cell lines LNCaP,

DU145 and PC3 to determine effective concentrations that would impact cell proliferation in a

measureable manner. We determined that low doses of docetaxel induced cell growth at .1

and 1nM concentrations. Low doses of cytotoxic drugs have been known to initiate oxidative

stress in cells, which acts as a mitogenic factor instead of inhibitory, as the cell easily

compensates for the low levels of stimulation, resulting in cell progression through the cell

cycle (Minelli et al. 2009). In all three cell lines 10nM and 100nM doses induced significant

decreases in cell proliferation and maximal decreases in proliferation respectively, suggesting

potential effective concentrations to utilise in combination with phenoxodiol.

Taxanes are used in the treatment of advanced prostate cancer and recurrent hormone

refractory and castrate resistant prostate cancer. Docetaxel, as with other taxanes, binds to

the β-tublin subunit in microtubulin, promotes polymerisation of tubulin and disrupts

microtubule dynamics. As a consequence of microtubulin stabilisation, cells become arrested

in G2/M phase and eventually undergo an apoptotic form of cell death. The effects of taxanes

may vary depending on cell type and drug concentration. These observations raised the

question as to what might be the intracellular signalling machinery that controls the apoptotic

differences in response to taxanes in prostate cancer. Studies have shown that apoptosis

induced by taxanes involves several apoptotic signal molecules, such as c-Jun N-terminal

Page 246: Investigating the effects of, the isoflavone, Phenoxodiol

226

kinase (JNK), protein kinase A (PKA), c-Raf-1/Ras/Bcl-2, p53/p21, and mitogen-activated

protein kinases ERK and p38 (Wang et al. 2008). Combined docetaxel and prednisone is

currently considered the standard of care for men with castrate resistant prostate cancer,

based largely on the simultaneous publication of Tannock and Petrylak (2004), where two large

randomised controlled trials compared this combination with the previously established

standard of mitoxantrone and prednisone. To determine an effective combination therapy

accurately, the interaction of two compounds is analysed and the effectiveness investigated

for synergistic, additive or even interference effects using an isobologram chart where a range

of effective doses is compared with each other (Zhao et al. 2004; Tallarida 2006).

In this study we investigated the effects of phenoxodiol combined with docetaxel treatment in

an isobologram format over 48 hours, with 24 different potential combinations per cell line.

We determined that the 5µM phenoxodiol concentration was inducing an interference effect

against the 100nM docetaxel concentration. This was the highest concentration of docetaxel

used and the concentration that exhibited the maximal effect of the compound after 48 hours

of treatment. All three cell lines had significant interference effects caused by phenoxodiol and

docetaxel combination at that concentration while the DU145 cell line also had the 10µM

phenoxodiol dose significantly impacting on the ability of docetaxel to induce cytotoxicity.

To further investigate these phenomena we evaluated pretreatment combinations in

conjunction with complete combinations over 48 hours. We investigated the effects of treating

the cells with vehicle control, 10µM phenoxodiol, 100nM docetaxel or combination 10µM

phenoxodiol/100nM docetaxel for 24 hours. Following this, the two extra sets of cells received

an addition of 10µM phenoxodiol; or 100nM docetaxel giving them a total of 48 hours of

100nM docetaxel and 24 hours of phenoxodiol or 48 hours 10µM phenoxodiol and 24 hours of

Page 247: Investigating the effects of, the isoflavone, Phenoxodiol

227

100nM docetaxel. This was also repeated for 30µM phenoxodiol concentrations to determine

if the increased cytotoxicity of phenoxodiol would overcome the limited interference that had

appeared in the isobolograms. The cell lines indicated that, at all phenoxodiol concentrations

after 48 hours of phenoxodiol and 24 hours of docetaxel treatment, there was no significant

different versus phenoxodiol by itself, other than PC3 10µM samples. This was followed by the

result that, at both 10µM and 30µM concentrations, there was a significant increase in

proliferation versus 100nM docetaxel alone, meaning that phenoxodiol had significantly

impacted the ability of docetaxel to induce cell arrest at the G2/M stage. While synergism and

additivity are the targets of drugs, especially those with different modes of action,

phenoxodiol’s exhibited ability to induce cell arrest at G1 and S phases of the cell cycle

prevents the cells from entering the G2/M region where docetaxel can be effective (Aguero et

al. 2010). Studies have also indicated that increased expression of p21WAF1 corresponds to an

inhibition of docetaxel activity via a p38 dependent signalling pathway and that cell cycle

inhibitors can actually protect the cells from taxane induced cell death during certain periods

(Canfield et al 2006; Gan et al 2011). A better understanding of how such mechanisms work at

the molecular level may have implications in the rational use of isoflavone/taxane based

chemotherapy.

The canonical Wnt/β-catenin signalling pathway promotes transcriptional activity through

activation of downstream target genes such as c-Myc and Cyclin-D1 driving cellular

proliferation (Menezes et al 2012; Watcharasit et al 2002). The Wnt ligand target, the frizzled

receptor, is overexpressed in prostate cancer but that this overexpression is counterbalanced

by the secreted frizzled related protein (sFRP) family which attempt to suppress AR-mediated

transactivation (Kawano et al. 2009). Methylation of the sFRP proteins has been suggested as a

marker of invasive carcinoma with a resulting poor prognosis (Ahmad et al. 2011) and

conversely, a total lack of β-catenin in a prostate cell has also been implicated in metastatic

Page 248: Investigating the effects of, the isoflavone, Phenoxodiol

228

formation (Aaltomaa et al. 2005). Unlike other inhibitors of Wnt signalling, sFRP4 appears to

affect androgen-dependent and androgen-independent prostate cancer (Horvath et al. 2007).

In this study we investigated the effects of purified sFRP4 protein in conjunction with

phenoxodiol, investigating the published abilities of isoflavones to interact with the Wnt

pathway (Li et al. 2008; Liss et al. 2010). Purified protein samples induced a decrease in cell

proliferation in LNCaP and D145 cells but not in PC3 cells. All cell lines exhibited a significantly

decreased proliferation rate when 500pg/mL purified protein was combined with 30µM

phenoxodiol over 48 hours of treatment, beyond that of 30µM phenoxodiol by itself. The

ability for sFRP4 to interact with frizzled results in a stabilisation of the GSK3β molecule, which

has been shown to be critical for soy based molecules, such as isoflavones, to induce cell

toxicity (Liss et al. 2010). This is one manner in which sFRP4 could be sensitising prostate

cancer cells to phenoxodiol treatment; others include the increased expression of p21WAF1 after

frizzled activation (Hall et al. 2010), GSK3β binding to and inactivating IAP members (Li et al.

2008), stabilisation of p53 expression (Watcharasit et al. 2002) or even downregulation of the

AR through GSK3β stabilisation. The Wnt pathway affects many parts of cell homeostasis and

the ability to prevent Wnt ligand activation of β-catenin and, therefore, stabilise GSK3β results

in a signalling situation in which phenoxodiol is more effective.

In this study we demonstrate that the cytotoxic taxane compound, docetaxel, induces cell

death that is attenuated by co-treatment or pre-treatment of cells with phenoxodiol. This

attenuation is associated with the prevention of cells from entering the G2/M phase of the cell

cycle where docetaxel is functional, damaging the spindle fibres and potentially due to p21WAF1

mediated cell survival after docetaxel treatment. Increased expression of p21WAF1, p53 and p38

has been determined to inhibit the ability of docetaxel to induce cytotoxicity (Gan et al. 2011).

Page 249: Investigating the effects of, the isoflavone, Phenoxodiol

229

We also exhibit the ability of sFRP4 protein to increase the effectiveness of phenoxodiol

treatment through alteration of the Wnt/β-catenin signalling pathway. Through stabilisation of

the GSK3β molecule, sFRP4 induces degradation of active β-catenin which causes an increased

sensitivity to isoflavone cytotoxic induction by increasing p21WAF1 expression and decreases

expression of c-Myc, Cyclin-D1 and other potent oncogenes. Phenoxodiol induces significant

cytotoxicity when combined with a Wnt/β-catenin receptor blocker such as sFRP4. This

promotes the concept that combination therapy of a Wnt inhibitor with phenoxodiol might

increase the effectiveness of phenoxodiol and give a subset population of prostate cancer

sufferers a more effective treatment regime.

Page 250: Investigating the effects of, the isoflavone, Phenoxodiol

230

8 . C H A P TER E IGH T: G EN ER A L D IS C US S ION

8.1. DISCUSSION

Phenoxodiol, [2H-1-Benzopyran-7-0,1,3-(4-hydroxyphenyl)], is an isoflavone derivative that has

also been shown to elicit cytotoxic effects against a broad range of human cancers. Currently

undergoing human clinical trials, it has shown promise in patients with recurrent ovarian

cancer where the cancer is refractory or resistant to standard chemotherapy, and in patients

with hormone-refractory prostate cancer (Sapi et al. 2004; Brown et al. 2005). Preliminary

studies involving a number of flavanoid derivatives have demonstrated that phenoxodiol

inhibits cell proliferation in a wide range of human cancer cell lines including leukaemia, breast

and prostate carcinomas, and is five to twenty times more potent than a similar compound,

Genistein (Aguero et al. 2005). Phenoxodiol has been characterised in ovarian cancer as

affecting key ovarian anti-apoptotic signalling pathways (Kamsteeg et al. 2003) as well as

reversing the ability of cells to become resistant to Docetaxel, through over expression of anti-

apoptotic molecules (Sapi et al. 2004). In breast cancer phenoxodiol acts as an inhibitor of cell

division (Constantinou et al. 2003), whilst in prostate cancer cells co-cultured with osteoblasts

phenoxodiol downregulated the cancer specific enzyme tNADH-Oxidase (Axanova et al. 2005).

Finally, it has been demonstrated that phenoxodiol induces G1 specific arrest through loss of

Cyclin-Dependant Kinase 2 activity by p53-independent induction of p21WAF1 in a battery of

human cell lines and that phenoxodiol affects multiple cancer types through prevention of cell

replication (Aguero et al. 2005; Aguero et al. 2010).

The ability to induce high rates of cytotoxicity in castrate resistant prostate cancer cells is the

goal of mainstream treatment regimes. Investigating the cytotoxicity of phenoxodiol on

Page 251: Investigating the effects of, the isoflavone, Phenoxodiol

231

prostate cancer cells is driven by the historically low 5-year survival rate for advanced prostate

cancer where chemotherapy is palliative in nature (Kessler and Albertsen 2003). In this study

we initially sought to characterise the effects of phenoxodiol on the prostate cancer cell lines

LNCaP (AR positive/p53 wild type), DU145 (AR negative p53-mutant) and PC3 (AR negative

p53-null). These cell lines represent early to late stage prostate cancers phenotypes and exhbit

altered expression of p53, a commonly mutated gene. We compared the action of phenoxodiol

against molecules regulating some of the key hallmarks of cancer stated by (Hanahan and

Weinberg 2000) and then updated by (Trosko et al. 2004). We targeted a framework of genes

that regulated the cell’s ability to be; self-sufficient in growth signals; insensitive to anti-growth

signals; able to evade apoptosis and able to have limitless replicative potential.

Determining appropriate initial cell concentration rates ensured logarithmic growth

throughout the study and ensured that phenoxodiol elicits time- and dose-dependent anti-

proliferative activity against both androgen-responsive and androgen-resistant prostate cancer

cell lines. LNCaP cells were found to be less sensitive to phenoxodiol induced cytotoxicity

measured by the MTS proliferation assay than visual analysis suggested. (Wang et al. 2010)

reported an underestimating of anti-proliferative effects of cytotoxic compounds due to MTS

dye in LNCaP cells when compared to other techniques. The use of apoptotic assays 3’-end

labelling, JC-1 and AV/PI flow cytometry ensured accurate cytotoxicity information was

obtained. LNCaP, DU145 and PC3 cells exhibited significant increases in mitochondrial

depolarisation, causing the release of Cytochrome c, Bax and other potent pro-apoptotic

molecules into the cytoplasm, where they could induce cell cytotoxicity. LNCaP cells had the

largest mitochondrial depolarisation response, appropriate for an early stage prostate cancer

model, while DU145 and PC3 cells had less total depolarisation, fitting their characterisation as

chemoresistant late stage prostate cancer cell representatives. The flow cytometry and 3’-end

labelling study determined that LNCaP and DU145 responded apoptotically to treatment while

Page 252: Investigating the effects of, the isoflavone, Phenoxodiol

232

PC3 cells responded necrotically. Cytotoxicity was determined to not be induced by pH

changes in the media.

Classic apoptosis, such as that induced by chemotherapeutic agents, can proceed via extrinsic

death receptor-mediated and intrinsic mitochondrial-mediated pathways ultimately resulting

in the activation of Caspase-3 (Asselin et al. 2001). Phenoxodiol has been shown to induce

classic apoptosis in ovarian cancer (Gamble et al. 2005; Alvero et al. 2006). However, this

current study did not detect activated Caspase-3 expression increase in LNCaP and PC3 cells,

while the actual increase in expression of Caspase-3 in DU145 cells was minimal. The lack of

consistent signalling indicated that cell cytotoxicity induced by phenoxodiol was potentially

initiated via a caspase independent cell death pathway. Phenoxodiol exhibited significant

cytotoxicity, inducing cell death in the prostate cancer cell lines LNCaP, DU145 and PC3 by

apoptotic and necrotic responses. All three cell lines demonstrated significantly decreased

viable cell populations after only 48 hours of treatment as determined by the AV/PI FACS

analysis technique, the JC-1 mitochondrial depolarisation assay indicating a method of

apoptotic induction via intrinsic cytotoxic signalling.

Two of the hallmarks of cancer initiation and progression are the ability of cells to become

insensitive to anti-growth signals and to evade programmed cell death, which is caused by

accumulated increased expression of anti-apoptotic signalling molecules that prevent

activation of the intrinsic and extrinsic signalling pathways that normally induce cell death

(Hanahan and Weinberg 2000). We reported that phenoxodiol does not induce cytotoxicity

through direct manipulation of the expression of key elements of the extrinsic and intrinsic cell

death pathways. Though phenoxodiol induced mitochondrial depolarisation and a significant

Page 253: Investigating the effects of, the isoflavone, Phenoxodiol

233

decrease in cell population, expression of the pro-apoptotic AIF, Bax, Caspase-3 and anti-

apoptotic Bcl-xL, xIAP were not consistently altered after phenoxodiol treatment.

A key aspect of chemotherapeutic treatments is the ability to induce cytotoxicity reliably over

a range of cells representing multiple stages of development, with a method of action that can

be consistently determined, as this maximises potential treatment candidates and successful

application. Though we detected apoptosis and mitochondrial depolarisation, we determined

that phenoxodiol acted in a caspase-independent manner. When phenoxodiol was combined

with a broad spectrum caspase inhibitor there was no effect on the ability of phenoxodiol to

induce cell death. In this study we determined that the cytotoxicity of phenoxodiol was not

associated with consistent changes in both pro-apoptotic and anti-apoptotic members of the

Bcl-2 family between the cell lines, unlike previous studies (Aguero et al. 2005; Yu et al. 2006).

LNCaP displayed a decreased expression of Bax protein and increased expression of Bcl-xL

protein, with cells attempting to prevent the effect of phenoxodiol-induced mitochondrial

depolarisation by mediating the effects of the Bcl-2 family. However, LNCaP cells still exhibited

significant cytotoxicity to treatment, indicating that the classic intrinsic apoptotic pathway

might not be directly responsible for the effect of phenoxodiol treatment. DU145 cells did not

express Bax protein at all, influencing the cell into a strong anti-apoptotic phenotype; yet a

detectable decrease in Bcl-xL protein was determined after treatment indicating a

susceptibility to phenoxodiol-induced mitochondrial depolarisation with resultant

mitochondrial efflux. This is in conjunction with PC3 cells, whose levels of Bax were unchanged

due to treatment but Bcl-xL levels were significantly reduced, indicating that the ratio of Bcl-

xL:Bax had significantly decreased; a sign of cell death signalling sensitivity from this

notoriously chemoresistant cell line (Li et al. 2008).

The lack of a consistent change in gene expression between LNCaP, DU145 and PC3 cell lines

indicates that phenoxodiol does not promote cell death by disrupting the stoichiometry of key

Page 254: Investigating the effects of, the isoflavone, Phenoxodiol

234

pro- and anti-apoptotic molecules studied here in prostate cancer cells. The common model

for cancer therapy activity has been that anti-cancer regimens cause apoptotic signalling to

occur, often mediated by p53 and Caspase-3 upregulation (Fojo and Bates 2003). Therefore,

cells that are resistant to apoptosis achieve this through forced over expression of anti-

apoptotic molecules such as xIAP and Bcl-xL family members or p53 mutation (Brown and

Attardi 2005). Members of the Inhibitors of Apoptosis Proteins (IAP), such as xIAP, can prevent

apoptosis by inactivating Caspase-3 and Caspase-9. The results reported in this study are in

contrast to other reports where it has been demonstrated in ovarian cancer that caspase

activation, Bax upregulation and xIAP degradation are key proteins modulated by phenoxodiol

(Sapi et al. 2004; Alvero et al. 2006; Mor et al. 2006; Kluger et al. 2007). The inability of the pan

caspase inhibitor, Z-VAD-FMK, to prevent the anti-proliferative activity of phenoxodiol in all

three cell lines reinforces that the phenoxodiol-induced mechanism of cell death in prostate

cancer cell lines is occurring via a caspase-independent pathway. This is in contrast to many

studies that indicate PARP and caspase activation as being the effect of phenoxodiol treatment

(Kamsteeg et al. 2003; Aguero et al. 2005; Aguero et al. 2010).

One potential method of cytotoxic induction investigated was the ability of phenoxodiol to

induce accumulation/production of reactive oxygen species (ROS) in the cell, causing oxidative

stress. Oxidative stress has been linked to mitochondrial depolarisation but we determined

that, in all cell lines, phenoxodiol did not induce nitric oxide production as a response to

treatment, even though all three cell lines retained the ability to have NO production

stimulated by DEAN and SNP addition.

In addition to apoptosis, cancer cells can be effectively eliminated through necrosis, mitotic

catastrophe and premature senescence, which result in cell cycle arrest and subsequent cell

Page 255: Investigating the effects of, the isoflavone, Phenoxodiol

235

death signalling (Brown and Wouters 1999; Tannock and Lee 2001; Brown and Wilson 2003).

Cells that are non-responsive to treatment through the classic intrinsic apoptosis pathways,

yet have an apoptotic response, must activate this process through other mechanisms such as

mitochondrial response and depolarisation, mitotic catastrophe and DNA degradation (Ruth

and Roninson 2000; Brown and Wilson 2003). Two of the hallmarks of cancer are the ability to

be self-sufficient in growth signals and to have infinite replicative potential, which can be

initiated through damaging the cell cycle restriction points, thereby allowing for progression of

cancer from a benign to metastatic state (Hanahan and Weinberg 2000). This study indicates

that phenoxodiol induces cell cycle arrest in the G1 and S phase of the cell cycle, with the

resultant arrest due to the upregulation of p21WAF1 and cytotoxicity due to resultant

downstream apoptotic signalling events. c-Myc and Cyclin-D1 expression was not consistently

altered in response to phenoxodiol treatment but Ki-67 was down regulated in LNCaP and PC3

cells and, though not significant, there was a similar trend in the DU145 cells.

The tumour suppressor p53 is the primary controller of cell cycle activity, which triggers cell

cycle arrest; but is only active in LNCaP cells with DU145 cells expressing a mutated p53 and

PC3 cells being p53-null. p21WAF1 is a cyclin dependant kinase inhibitor (CdkI) family member,

along with p27 and p57, which interfere with the cyclin dependant kinase-cyclin complex. The

ability of isoflavones, and specifically phenoxodiol, to induce cell cycle arrest has been

previously reported with studies indicating that arrest was induced by p21WAF1 stabilisation and

expression increase (Aguero et al. 2005; Aguero et al. 2010; Seo et al. 2011). In this study we

exhibited the ability of phenoxodiol to increase p21WAF1 expression, with the resultant cell

cycle arrest at both 10µM and 30µM concentrations over 24 and 48 hours of treatment. The

cell lines exhibited significantly decreased populations of G2 cells and arrest of the cell cycle,

visible in the significant alteration of G1 and S phase cell populations. It’s clear that the

Page 256: Investigating the effects of, the isoflavone, Phenoxodiol

236

method of phenoxodiol-induced cell cycle arrest is independent of p53 status and caspase

activity.

The canonical Wnt/β-catenin signalling pathway controls cellular interaction with growth

signals, where Wnt signalling causes decreased expression of GSK3β protein which results in an

active form of β-catenin translocating to the nucleus and initiating cell cycle progression

through multiple pathways including Cyclin-D1 and c-Myc expression (Zhang et al. 2011). We

investigated the expression of the active form of β-catenin and determined that LNCaP and

PC3 cells had significantly reduced active β-catenin expression after 24 hours, with LNCaP

continuing on to a significant 48 hour effect, which coincides with a subsequent decrease in

Cyclin-D1 signalling expression in androgen insensitive PC3 cells, while LNCaP expression is

unchanged due to the potential moderation of the AR receptor. DU145 cells exhibited an

increased expression of active β-catenin after treatment over 24 and 48 hours, indicating that

alteration of canonical Wnt signalling is not a direct effect of phenoxodiol treatment but a

downstream signalling outcome.

Secreted frizzled related protein 4 (sFRP4) is one of a group of proteins that antagonise the

canonical Wnt/β-catenin pathway, decreasing Wnt activity and causing decreased activated β-

catenin formation and decreased invasive potential. We determined that sFRP4 expression

was increased over 24 hours in PC3 cells coinciding with decreased active β-catenin expression

in that cell line. sFRP4 expression increased over 24 and 48 hours in LNCaP cells, coinciding

with decreased active β-catenin in that cell line. sFRP4 expression was increased in DU145 cells

but did not correspond with β-catenin expression changes, indicating that DU145 cells might

have an modification of the canonical Wnt/β-catenin signalling pathway. Separate from β-

catenin degradation it has been suggested that sFRP4 inhibition of Wnt signalling causes

Page 257: Investigating the effects of, the isoflavone, Phenoxodiol

237

stabilisation of GSK3β resulting in p53 activation and associated signalling such as p21WAF1, as

well as a decreased invasive potential in androgen-independent prostate cancer cells.

Addiction of purified sFRP4 protein resulted in a decrease in cell proliferation in LNCaP and

D145 cells but not in PC3 cells. All cell lines exhibited a significantly decreased proliferation

rate when 500pg/mL purified protein was combined with 30µM phenoxodiol over 48 hours of

treatment, beyond that of 30µM phenoxodiol by itself. The ability for purified sFRP4 to interact

with frizzled results in a potential stabilisation of the GSK3β molecule and p21WAF1 expression

alteration which (Liss et al. 2010) has shown to be critical for soy based molecules, like

isoflavones, to induce cell toxicity.

Docetaxel and paclitaxel are from a class of anti-cancer agents called taxanes that bind to and

stabilise microtubules, causing G2/M cell-cycle arrest and cell death (Montero et al. 2005). In

this study we investigated the effects of docetaxel on the prostate cancer cell lines LNCaP,

DU145 and PC3 to determine effective concentrations that would impact cell proliferation in a

measureable manner. In all three cell lines 10nM and 100nM doses induced significant

decreases in cell proliferation and maximal decreases in proliferation respectively, suggesting

potential effective concentrations to utilise in combination with phenoxodiol.

Taxanes are used in the treatment of advanced prostate cancer and recurrent, hormone

refractory and castrate-resistant prostate cancer. Docetaxel, as with other taxanes, binds to

the β-tublin subunit in microtubulin, promotes polymerisation of tublin and disrupts

microtubule dynamics. As a consequence of microtubulin stabilisation, cells become arrested

in G2/M phase and eventually undergo an apoptotic form of cell death. In this study we

investigated the effects of phenoxodiol combined with docetaxel treatment in an isobologram

format over 48 hours, with 24 different potential combinations per cell line. We determined

Page 258: Investigating the effects of, the isoflavone, Phenoxodiol

238

that the 5µM phenoxodiol concentration was inducing an interference effect against the

100nM docetaxel concentration. All three cell lines exhibited significant interference effects

caused by phenoxodiol and docetaxel combination at the specific combination, while DU145

also had the 10µM phenoxodiol dose impacting on the ability of docetaxel to induce

cytotoxicity.

Pretreatment drug combinations in conjunction with complete combinations over 48 hours

were then investigated. The cell lines indicated that, at all phenoxodiol concentrations after 48

hours of phenoxodiol and 24 hours of docetaxel treatment, there was no significant difference

in effect versus phenoxodiol by itself, other than PC3 10µM samples. This is followed by the

result that, at both 10 and 30µM concentrations, there was a significant increase in

proliferation versus 100nM docetaxel alone; indicating that phenoxodiol had significantly

impacted the ability of docetaxel to induce cell arrest at the G2/M stage. While synergism and

additivity are the targets of drugs, especially those with different modes of action,

phenoxodiol’s exhibited ability to induce arrest at G1 and S phases of the cell cycle prevents

the cells from entering the G2/M region, where docetaxel can be effective (Aguero et al.

2010). Studies have also indicated that increased expression of p21WAF1 corresponds to an

inhibition of docetaxel activity via a p38-dependent signalling pathway, and that cell cycle

inhibitors can actually protect the cells from taxane-induced cell death during certain periods

(Canfield et al. 2006; Gan et al. 2011). A better understanding of how such mechanisms work

at the molecular level may have implications in the rational use of isoflavone/taxane based

chemotherapy.

We have exhibited the ability of phenoxodiol to induce cytotoxicity and cell death through a

variety of means in a p53 and caspase-independent manner by arrest of cells in the G1 and S

Page 259: Investigating the effects of, the isoflavone, Phenoxodiol

239

phases of the cell cycle, thereby preventing cells from moving into G2/M phase. The cell

cytotoxicity was partly mediated by mitochondrial efflux, with all cells responding with a

degree of total efflux of factors into the cytoplasm in response to phenoxodiol treatment. The

cell cycle arrest was mediated by expression increases in p21WAF1 and the resultant cytotoxic

induction resulted in an apoptotic and necrotic response, independent of p53 status. We also

demonstrated that pre-treatment or co-treatment of prostate cancer cells with docetaxel and

phenoxodiol actually caused an interference response; and that co-treatment with sFRP4

purified protein and phenoxodiol exhibited a sensitisation effect, potentially caused by the

stabilisation of the GSK3β molecule and a subsequent increase in p21WAF1 signalling.

8.2. CONCLUSION

We investigated the ability of Phenoxodiol to induce cytotoxicity in prostate cancer cell lines

representing early and late stage prostate cancer. We then determined the extent of induced

cytotoxicity by using multiple qualitative and quantitative measures to characterise the mode

of action. We can conclude that phenoxodiol induces significant cellular cytotoxicity over 24

and 48 hours of treatment with 10µM and 30µM doses of phenoxodiol. We can also state that

phenoxodiol induces apoptosis in LNCaP and DU145 cell lines while PC3 cell lines respond

necrotically, as evidenced by apoptotic DNA fragmentation in LNCaP and DU145 cells and DNA

smearing in PC3, similarly confirmed by AV/PI flow cytometry. We also determined that

Caspase-3 expression is not significantly up regulated in response to phenoxodiol treatment

except in DU145 cells; indicating that phenoxodiol’s method of action in prostate cancer cells

might be caspase independent.

Page 260: Investigating the effects of, the isoflavone, Phenoxodiol

240

We investigated the underlying apoptotic signalling machinery, after phenoxodiol treatment,

examining the link between early and late stage apoptotic signalling molecule expression. We

then determined that phenoxodiol had the ability to induce cytotoxicity through means other

than the direct innervation of intrinsic and extrinsic caspase signalling pathways. The analysis

of the anti-apoptotic Bcl-xL and xIAP molecules and pro-apoptotic AIF, Bax and Caspase-3

molecules did not show evidence of consistent signalling response as a result of phenoxodiol

treatment. Following this, a pan caspase inhibition study indicated that caspase expression was

not critical for phenoxodiol’s method of action; in contrast to previously published data and

important when many cancers have strong anti-apoptotic protein expression inhibiting

signalling through classic extrinsic and intrinsic apoptosis pathways.

We investigated the link between phenoxodiol and the cell cycle by determining the potential

effects on the various prostate cancer cell lines and looking for a common signalling pathway

in response to treatment. We also investigated the response of active β-catenin and sFRP4 to

phenoxodiol treatment after studies concluded evidence of a link between β-catenin

expression and cell cycle in prostate cancer. We determined that phenoxodiol prevented cell

cycle progression in a dose/time dependant manner through arrest in the G1 and S phase of

the cell cycle. This corresponded to an increase in P21WAF1 signalling in all cell lines. Alterations

in c-Myc, Cyclin-D1 and Ki-67 expression were detected, with only Ki-67 indicating a consistent

response. The expression of active β-catenin was down regulated in LNCaP and PC3 cells but

increased in DU145, indicating that phenoxodiol was not directly targeting β-catenin

expression in the canonical Wnt signalling pathway. However, sFRP4 levels were altered in a

manner that was indicative of the activated β-catenin decreases demonstrated. Both LNCaP

and PC3 cells exhibited increased expression of sFRP4 and resultant decreased expression of

active β-catenin protein.

Page 261: Investigating the effects of, the isoflavone, Phenoxodiol

241

We investigated the effects of combination therapy of phenoxodiol and docetaxel looking at

the effective dose of docetaxel, an isobologram of dose/response curves for phenoxodiol and

docetaxel therapy and the use of each as a pre-treatment for sensitisation. We also

investigated the effects of sFRP4 treatment in conjunction with phenoxodiol, targeting the cell

cycle through multiple means of interaction. We determined the docetaxel effective treatment

dose and, at that concentration, phenoxodiol treatment inhibits docetaxel at certain

combinations and concentrations by preventing cells from entering the G2/M phase of the cell

cycle where docetaxel exhibits its cell cycle arrest response. While high rates of cytotoxicity

were evident, we determined that combination therapy of docetaxel and phenoxodiol can

result in an interference reaction as expected. Phenoxodiol is effective at preventing cells from

progressing through G1/S phase, where docetaxel is effective, thereby preventing its ability to

induce G2 cell cycle arrest. Combination therapy with sFRP4 indicated a sensitisation to

phenoxodiol treatment, which was not necessarily synergistic, because sFRP4 didn’t exert a

consistent cytotoxic effect over 48 hours. However, interference with the canonical Wnt/β-

catenin pathway assisted the action of phenoxodiol. Therefore, Wnt/β-catenin antagonism

results in stabilisation of GSK3β which assists in the expression of p21WAF1.

In this thesis we provide evidence that the soy based isoflavone, phenoxodiol has the ability to

potentiate cell cytotoxicity through a variety of means including canonical Wnt/β-catenin

signalling interference, Cdk inhibitor p21WAF1 activation, cell cycle arrest independent of p53

status and caspase independent induction of apoptosis. We exhibited the ability of

phenoxodiol to induce significant cytotoxicity in cells, representing early and late stage

prostate cancer, over 24 and 48 hours of treatment with 10µM and 30µM concentrations.

Phenoxodiol activity was determined to be increased when applied in combination therapy

with a Wnt antagonist, sFRP4. This implicates sensitisation of tumours towards phenoxodiol-

induced cytotoxicity through antagonism of the Wnt receptor frizzled. Coupled with the

Page 262: Investigating the effects of, the isoflavone, Phenoxodiol

242

reported ability of high tolerance of orally ingested phenoxodiol, and few reported side

effects, Phenoxodiol represents a strong, effective, potential treatment for all stages of

prostate cancer including the currently incurable castrate-resistant late stage prostate cancer.

8.3. LIMITATIONS

My perceived limitations of the study are as follows. Several grams only of phenoxodiol was

provided once for the entirety of the thesis experimental use. As such the IC50 concentration

was not determined, merely a response to treatment was initially sought and the use of 10µM

and 30µM concentrations was suggested by Novogen Pty Ltd as doses reaching biological

efficacy. The doses were kept at these concentrations due to the 2 week lifespan of

phenoxodiol once diluted in DMSO and higher concentrations would result in a rapid depletion

of the compound. Docetaxel also suffered from a short lifespan once diluted but was available

for purchase once supplies were diminished whereas phenoxodiol was not.

My cell cycle data was completed and provided to Novogen 18 months prior to the publishing

of the 2010 Aguero article which contains data very similar to what I found. Hence I view this

thesis to be a complete work of, at the time of experimental completion, unknown responses

to phenoxodiol as I was not informed about any data produced by other research groups being

provided with phenoxodiol. With the published article of 2010 indicating that PXD affected cell

cycle and P21WAF1 status in prostate cancer cells, we can independently confirm their findings

while also noting that we determined significant increases in S phase and G1 phase arrest.

In vivo experimentation in Nude mice was explored but determined to be difficult to complete

effectively with all three cell lines represented. The use of a 3D ultrasound capable of mapping

Page 263: Investigating the effects of, the isoflavone, Phenoxodiol

243

tumour volume and blood flow was not available within the state and getting access was cost

prohibitive. Animal ethics raised concerns over the use of multiple anaesthesia to measure

tumour growth and regression and thought a larger n value would be more appropriate but

lacking in the ability to track tumour changes over a course of doses.

Silencing work would have been an effective method to deduce the importance of p21WAF1

expression to phenoxodiol activity however the reported p27 and p16 pathway upregulation

and unknown conservation status of p21 WAF1 in various forms of cancer has since indicated

that p21 is not the sole target of phenoxodiol activity and silencing it would not necessarily

have resulted in an inability for phenoxodiol to induce cytotoxicity through cell cycle arrest.

The decreased sensitivity of LNCaP to the MTS proliferation assay, as described by (Wang et al.

2010), was problematic and prevented it as a sole measure of effective proliferation decrease

between cell lines, but still allowed for comparisons within the treatments of the LNCaP cell

line. The MTS assay was primarily utilised for its economy of scale that allowed the 3 cell lines

and multiple treatments to be effectively treated, stained and analysed within short periods of

each other. Other proliferation analysis methods were investigated but tritiated thymidine was

cost prohibitive and radioactive, XTT/MTT assays ran into exactly the same problem and were

more expensive, trypan blue viability staining would be excessively time intensive for the

purpose and the volume of cells/treatments being analysed in one point would have been

decreased. Cell density analysis was limited by access to the equipment which was granted and

then permission was retracted after 2 months of experiments meaning I couldn’t compare or

utilise the data generated by that technique.

Page 264: Investigating the effects of, the isoflavone, Phenoxodiol

244

9 . B IB L IOGRA P H Y

Aaltomaa, S., V. Karja, P. Lipponen, T. Isotalo, J. P. Kankkunen, M. Talja and R. Mokka (2005). "Reduced alpha- and beta-catenin expression predicts shortened survival in local prostate cancer." Anticancer Res 25(6C): 4707-4712.

Aguero, M. F., M. M. Facchinetti, Z. Sheleg and A. M. Senderowicz (2005). "Phenoxodiol, a novel isoflavone, induces G1 arrest by specific loss in cyclin-dependent kinase 2 activity by p53-independent induction of p21WAF1/CIP1." Cancer Res 65(8): 3364-3373.

Aguero, M. F., M. Venero, D. M. Brown, M. E. Smulson and L. A. Espinoza (2010). "Phenoxodiol inhibits growth of metastatic prostate cancer cells." Prostate 70(11): 1211-1221.

Ahmad, I., R. Patel, Y. Liu, L. B. Singh, M. M. Taketo, X. R. Wu, H. Y. Leung and O. J. Sansom (2011). "Ras mutation cooperates with beta-catenin activation to drive bladder tumourigenesis." Cell Death Dis 2: e124.

AIHW, Ed. (2010). Cancer in Australia 2010: an overview. Cancer Series. Canberra, AIHW.

Alhasan, S. A., H. Pietrasczkiwicz, M. D. Alonso, J. Ensley and F. H. Sarkar (1999). "Genistein-induced cell cycle arrest and apoptosis in a head and neck squamous cell carcinoma cell line." Nutr Cancer 34(1): 12-19.

Alimirah, F., J. Chen, Z. Basrawala, H. Xin and D. Choubey (2006). "DU-145 and PC-3 human prostate cancer cell lines express androgen receptor: implications for the androgen receptor functions and regulation." FEBS Lett 580(9): 2294-2300.

Alvero, A. B., D. Brown, M. Montagna, M. Matthews and G. Mor (2007). "Phenoxodiol-Topotecan Co-Administration Exhibit Significant Anti-Tumor Activity Without Major Adverse Side Effects." Cancer Biol Ther 6(4).

Alvero, A. B., M. Kelly, P. Rossi, A. Leiser, D. Brown, T. Rutherford and G. Mor (2008). "Anti-tumor activity of phenoxodiol: from bench to clinic." Future Oncol 4(4): 475-482.

Alvero, A. B., D. O'Malley, D. Brown, G. Kelly, M. Garg, W. Chen, T. Rutherford and G. Mor (2006). "Molecular mechanism of phenoxodiol-induced apoptosis in ovarian carcinoma cells." Cancer 106(3): 599-608.

Amantana, A., C. A. London, P. L. Iversen and G. R. Devi (2004). "X-linked inhibitor of apoptosis protein inhibition induces apoptosis and enhances chemotherapy sensitivity in human prostate cancer cells." Mol Cancer Ther 3(6): 699-707.

Arnold, J. T. and J. T. Isaacs (2002). "Mechanisms involved in the progression of androgen-independent prostate cancers: it is not only the cancer cell's fault." Endocr Relat Cancer 9(1): 61-73.

Asselin, E., G. B. Mills and B. K. Tsang (2001). "XIAP regulates Akt activity and caspase-3-dependent cleavage during cisplatin-induced apoptosis in human ovarian epithelial cancer cells." Cancer Res 61(5): 1862-1868.

Avraamides, C. J., B. Garmy-Susini and J. A. Varner (2008). "Integrins in angiogenesis and lymphangiogenesis." Nat Rev Cancer 8(8): 604-617.

Page 265: Investigating the effects of, the isoflavone, Phenoxodiol

245

Axanova, L., D. J. Morre and D. M. Morre (2005). "Growth of LNCaP cells in monoculture and coculture with osteoblasts and response to tNOX inhibitors." Cancer Lett 225(1): 35-40.

Baldin, V., J. Lukas, M. J. Marcote, M. Pagano and G. Draetta (1993). "Cyclin D1 is a nuclear protein required for cell cycle progression in G1." Genes Dev 7(5): 812-821.

Basmajian, J. and C. Slonecker (1989). Grant's Method of Anatomy: A clinical Problem-Solving Approach. Baltimore, Hong Kong, London, Sydney, Williams & Wilkins.

Beecher, G. R. (2003). "Overview of dietary flavonoids: nomenclature, occurrence and intake." J Nutr 133(10): 3248S-3254S.

Begg, S., . Vos, T,. Barker, B,. Stevenson, C,. Stanley, L,. Lopez, AD,. Ed. (2007). The burden of disease and injury in Australia 2003. Canberra, AIHW.

Berezovskaya, O., A. D. Schimmer, A. B. Glinskii, C. Pinilla, R. M. Hoffman, J. C. Reed and G. V. Glinsky (2005). "Increased expression of apoptosis inhibitor protein XIAP contributes to anoikis resistance of circulating human prostate cancer metastasis precursor cells." Cancer Res 65(6): 2378-2386.

Berry, W. and M. Eisenberger (2005). "Achieving treatment goals for hormone-refractory prostate cancer with chemotherapy." Oncologist 10 Suppl 3: 30-39.

Birkey Reffey, S., J. U. Wurthner, W. T. Parks, A. B. Roberts and C. S. Duckett (2001). "X-linked inhibitor of apoptosis protein functions as a cofactor in transforming growth factor-beta signaling." J Biol Chem 276(28): 26542-26549.

Boatright, K. M. and G. S. Salvesen (2003). "Mechanisms of caspase activation." Curr Opin Cell Biol 15(6): 725-731.

Bott, S. R., M. Arya, R. S. Kirby and M. Williamson (2005). "p21WAF1/CIP1 gene is inactivated in metastatic prostatic cancer cell lines by promoter methylation." Prostate Cancer Prostatic Dis 8(4): 321-326.

Brown, D. M., G. E. Kelly and A. J. Husband (2005). "Flavonoid compounds in maintenance of prostate health and prevention and treatment of cancer." Mol Biotechnol 30(3): 253-270.

Brown, J. M. and L. D. Attardi (2005). "The role of apoptosis in cancer development and treatment response." Nat Rev Cancer 5(3): 231-237.

Brown, J. M. and G. Wilson (2003). "Apoptosis genes and resistance to cancer therapy: what does the experimental and clinical data tell us?" Cancer Biol Ther 2(5): 477-490.

Brown, J. M. and B. G. Wouters (1999). "Apoptosis, p53, and tumor cell sensitivity to anticancer agents." Cancer Res 59(7): 1391-1399.

Brusselmans, K., R. Vrolix, G. Verhoeven and J. V. Swinnen (2005). "Induction of cancer cell apoptosis by flavonoids is associated with their ability to inhibit fatty acid synthase activity." J Biol Chem 280(7): 5636-5645.

Canfield, S. E., K. Zhu, S. A. Williams and D. J. McConkey (2006). "Bortezomib inhibits docetaxel-induced apoptosis via a p21-dependent mechanism in human prostate cancer cells." Mol Cancer Ther 5(8): 2043-2050.

Page 266: Investigating the effects of, the isoflavone, Phenoxodiol

246

Carmeliet, P. and R. K. Jain (2011). "Principles and mechanisms of vessel normalization for cancer and other angiogenic diseases." Nat Rev Drug Discov 10(6): 417-427.

Cassinelli, G., R. Supino, V. Zuco, C. Lanzi, A. I. Scovassi, S. C. Semple and F. Zunino (2004). "Role of c-myc protein in hormone refractory prostate carcinoma: cellular response to paclitaxel." Biochem Pharmacol 68(5): 923-931.

Cheng, E. H., M. C. Wei, S. Weiler, R. A. Flavell, T. W. Mak, T. Lindsten and S. J. Korsmeyer (2001). "BCL-2, BCL-X(L) sequester BH3 domain-only molecules preventing BAX- and BAK-mediated mitochondrial apoptosis." Mol Cell 8(3): 705-711.

Chesire, D. R., C. M. Ewing, W. R. Gage and W. B. Isaacs (2002). "In vitro evidence for complex modes of nuclear beta-catenin signaling during prostate growth and tumorigenesis." Oncogene 21(17): 2679-2694.

Chomczynski, P. and N. Sacchi (1987). "Single-step method of RNA isolation by acid guanidinium thiocyanate-phenol-chloroform extraction." Anal Biochem 162(1): 156-159.

Choueiri, T. K., R. Wesolowski and T. M. Mekhail (2006). "Phenoxodiol: isoflavone analog with antineoplastic activity." Curr Oncol Rep 8(2): 104-107.

Constantinou, A. I., R. Mehta and A. Husband (2003). "Phenoxodiol, a novel isoflavone derivative, inhibits dimethylbenz[a]anthracene (DMBA)-induced mammary carcinogenesis in female Sprague-Dawley rats." Eur J Cancer 39(7): 1012-1018.

Cooper, S. (2003). "Reappraisal of serum starvation, the restriction point, G0, and G1 phase arrest points." FASEB J 17(3): 333-340.

Cory, S. and J. M. Adams (2002). "The Bcl2 family: regulators of the cellular life-or-death switch." Nat Rev Cancer 2(9): 647-656.

Cregan, S. P., V. L. Dawson and R. S. Slack (2004). "Role of AIF in caspase-dependent and caspase-independent cell death." Oncogene 23(16): 2785-2796.

D'Antonio, J. M., C. Ma, F. A. Monzon and B. R. Pflug (2008). "Longitudinal analysis of androgen deprivation of prostate cancer cells identifies pathways to androgen independence." Prostate 68(7): 698-714.

Datta, B. and R. Datta (1999). "Induction of apoptosis due to lowering the level of eukaryotic initiation factor 2-associated protein, p67, from mammalian cells by antisense approach." Exp Cell Res 246(2): 376-383.

Davies, R., Frydenberg, M., Tulloch, A., Kelly, G. (2005). Final Results of a PhaseIb/IIa Study of Oral Phenoxodiol in Patients with Late-stage, Hormone-Refractory Prostate Cancer. American Association for Cancer Research. Anaheim, CA.: 1.

De Luca, T., E. Bosneaga, D. M. Morre and D. J. Morre (2008). "Downstream targets of altered sphingolipid metabolism in response to inhibition of ENOX2 by phenoxodiol." Biofactors 34(3): 253-260.

De Luca, T., D. M. Morre, H. Zhao and D. J. Morre (2005). "NAD+/NADH and/or CoQ/CoQH2 ratios from plasma membrane electron transport may determine ceramide and sphingosine-1-phosphate levels accompanying G1 arrest and apoptosis." Biofactors 25(1-4): 43-60.

Page 267: Investigating the effects of, the isoflavone, Phenoxodiol

247

de Souza, P. L., W. Liauw, M. Links, S. Pirabhahar, G. Kelly and L. G. Howes (2006). "Phase I and pharmacokinetic study of weekly NV06 (Phenoxodiol), a novel isoflav-3-ene, in patients with advanced cancer." Cancer Chemother Pharmacol 58(4): 427-433.

de Souza, P. L., P. J. Russell, J. H. Kearsley and L. G. Howes (2010). "Clinical pharmacology of isoflavones and its relevance for potential prevention of prostate cancer." Nutr Rev 68(9): 542-555.

Deveraux, Q. L., E. Leo, H. R. Stennicke, K. Welsh, G. S. Salvesen and J. C. Reed (1999). "Cleavage of human inhibitor of apoptosis protein XIAP results in fragments with distinct specificities for caspases." Embo J 18(19): 5242-5251.

Deveraux, Q. L., R. Takahashi, G. S. Salvesen and J. C. Reed (1997). "X-linked IAP is a direct inhibitor of cell-death proteases." Nature 388(6639): 300-304.

Devlin, H. L., P. C. Mack, R. A. Burich, P. H. Gumerlock, H. J. Kung, M. Mudryj and R. W. deVere White (2008). "Impairment of the DNA repair and growth arrest pathways by p53R2 silencing enhances DNA damage-induced apoptosis in a p53-dependent manner in prostate cancer cells." Mol Cancer Res 6(5): 808-818.

Dharmarajan, A. M., S. Hisheh, B. Singh, S. Parkinson, K. I. Tilly and J. L. Tilly (1999). "Antioxidants mimic the ability of chorionic gonadotropin to suppress apoptosis in the rabbit corpus luteum in vitro: a novel role for superoxide dismutase in regulating bax expression." Endocrinology 140(6): 2555-2561.

Dohi, T., K. Okada, F. Xia, C. E. Wilford, T. Samuel, K. Welsh, H. Marusawa, H. Zou, R. Armstrong, S. Matsuzawa, G. S. Salvesen, J. C. Reed and D. C. Altieri (2004). "An IAP-IAP complex inhibits apoptosis." J Biol Chem 279(33): 34087-34090.

Donepudi, M. and M. G. Grutter (2002). "Structure and zymogen activation of caspases." Biophys Chem 101-102: 145-153.

Drake, J., A. M. Shearwood, J. White, R. Friis, N. Zeps, A. Charles and A. Dharmarajan (2009). "Expression of secreted frizzled-related protein 4 (SFRP4) in primary serous ovarian tumours." Eur J Gynaecol Oncol 30(2): 133-141.

Drake, J. M., R. R. Friis and A. M. Dharmarajan (2003). "The role of sFRP4, a secreted frizzled-related protein, in ovulation." Apoptosis 8(4): 389-397.

Dubrez-Daloz, L., A. Dupoux and J. Cartier (2008). "IAPs: more than just inhibitors of apoptosis proteins." Cell Cycle 7(8): 1036-1046.

Dudderidge, T. J., S. R. McCracken, M. Loddo, T. R. Fanshawe, J. D. Kelly, D. E. Neal, H. Y. Leung, G. H. Williams and K. Stoeber (2007). "Mitogenic growth signalling, DNA replication licensing, and survival are linked in prostate cancer." Br J Cancer.

Edinger, A. L. and C. B. Thompson (2004). "Death by design: apoptosis, necrosis and autophagy." Curr Opin Cell Biol 16(6): 663-669.

el-Deiry, W. S., T. Tokino, V. E. Velculescu, D. B. Levy, R. Parsons, J. M. Trent, D. Lin, W. E. Mercer, K. W. Kinzler and B. Vogelstein (1993). "WAF1, a potential mediator of p53 tumor suppression." Cell 75(4): 817-825.

Page 268: Investigating the effects of, the isoflavone, Phenoxodiol

248

Farooqi, A. A., S. Mukhtar, A. M. Riaz, S. Waseem, S. Minhaj, B. A. Dilawar, B. A. Malik, A. Nawaz and S. Bhatti (2011). "Wnt and SHH in prostate cancer: trouble mongers occupy the TRAIL towards apoptosis." Cell Prolif 44(6): 508-515.

Feldman, B. J. and D. Feldman (2001). "The development of androgen-independent prostate cancer." Nat Rev Cancer 1(1): 34-45.

Fojo, T. and S. Bates (2003). "Strategies for reversing drug resistance." Oncogene 22(47): 7512-7523.

Fox, S. and A. Dharmarajan (2006). "WNT signaling in malignant mesothelioma." Front Biosci 11: 2106-2112.

Fu, M., C. Wang, Z. Li, T. Sakamaki and R. G. Pestell (2004). "Minireview: Cyclin D1: normal and abnormal functions." Endocrinology 145(12): 5439-5447.

Gamble, J. R., P. Xia, C. N. Hahn, J. J. Drew, C. J. Drogemuller, D. Brown and M. A. Vadas (2005). "Phenoxodiol, an experimental anticancer drug, shows potent antiangiogenic properties in addition to its antitumour effects." Int J Cancer.

Gan, L., J. Wang, H. Xu and X. Yang (2011). "Resistance to docetaxel-induced apoptosis in prostate cancer cells by p38/p53/p21 signaling." Prostate.

Gao, J., J. T. Arnold and J. T. Isaacs (2001). "Conversion from a paracrine to an autocrine mechanism of androgen-stimulated growth during malignant transformation of prostatic epithelial cells." Cancer Res 61(13): 5038-5044.

Georgaki, S., M. Skopeliti, M. Tsiatas, K. A. Nicolaou, K. Ioannou, A. Husband, A. Bamias, M. A. Dimopoulos, A. I. Constantinou and O. E. Tsitsilonis (2009). "Phenoxodiol, an anticancer isoflavene, induces immunomodulatory effects in vitro and in vivo." J Cell Mol Med.

Gleason, D. (1977). The Veteran's Administration Cooperative Urologic Research Group: histologic grading and clinical staging of prostatic carcinoma. Urologic Pathology: The Prostate. M. Tannenbaum. Philadelphia, Lea and Febiger: 171-198.

Gleave, M., H. Miyake and K. Chi (2005). "Beyond simple castration: targeting the molecular basis of treatment resistance in advanced prostate cancer." Cancer Chemother Pharmacol 56 Suppl 1: 47-57.

Goodman, S. B., K. Kugu, S. H. Chen, S. Preutthipan, K. I. Tilly, J. L. Tilly and A. M. Dharmarajan (1998). "Estradiol-mediated suppression of apoptosis in the rabbit corpus luteum is associated with a shift in expression of bcl-2 family members favoring cellular survival." Biol Reprod 59(4): 820-827.

Gross, A., J. M. McDonnell and S. J. Korsmeyer (1999). "BCL-2 family members and the mitochondria in apoptosis." Genes Dev 13(15): 1899-1911.

Hall, C. L., S. Kang, O. A. MacDougald and E. T. Keller (2006). "Role of Wnts in prostate cancer bone metastases." J Cell Biochem 97(4): 661-672.

Hall, C. L., H. Zhang, S. Baile, M. Ljungman, S. Kuhstoss and E. T. Keller (2010). "p21CIP-1/WAF-1 induction is required to inhibit prostate cancer growth elicited by deficient expression of the Wnt inhibitor Dickkopf-1." Cancer Res 70(23): 9916-9926.

Page 269: Investigating the effects of, the isoflavone, Phenoxodiol

249

Hamzah, J., M. Jugold, F. Kiessling, P. Rigby, M. Manzur, H. H. Marti, T. Rabie, S. Kaden, H. J. Grone, G. J. Hammerling, B. Arnold and R. Ganss (2008). "Vascular normalization in Rgs5-deficient tumours promotes immune destruction." Nature 453(7193): 410-414.

Hanahan, D. and R. A. Weinberg (2000). "The hallmarks of cancer." Cell 100(1): 57-70.

Harris, C. C. (1996). "Structure and function of the p53 tumor suppressor gene: clues for rational cancer therapeutic strategies." J Natl Cancer Inst 88(20): 1442-1455.

Hayflick, L. and P. S. Moorhead (1961). "The serial cultivation of human diploid cell strains." Exp Cell Res 25: 585-621.

He, T. C., A. B. Sparks, C. Rago, H. Hermeking, L. Zawel, L. T. da Costa, P. J. Morin, B. Vogelstein and K. W. Kinzler (1998). "Identification of c-MYC as a target of the APC pathway." Science 281(5382): 1509-1512.

Hedlund, T. E., W. U. Johannes and G. J. Miller (2003). "Soy isoflavonoid equol modulates the growth of benign and malignant prostatic epithelial cells in vitro." Prostate 54(1): 68-78.

Hempstock, J., J. P. Kavanagh and N. J. George (1998). "Growth inhibition of prostate cell lines in vitro by phyto-oestrogens." Br J Urol 82(4): 560-563.

Hengartner, M. O. (2000). "The biochemistry of apoptosis." Nature 407(6805): 770-776.

Herst, P. M., J. E. Davis, P. Neeson, M. V. Berridge and D. S. Ritchie (2009). "The anti-cancer drug, phenoxodiol, kills primary myeloid and lymphoid leukemic blasts and rapidly proliferating T cells." Haematologica 94(7): 928-934.

Herst, P. M., T. Petersen, P. Jerram, J. Baty and M. V. Berridge (2007). "The antiproliferative effects of phenoxodiol are associated with inhibition of plasma membrane electron transport in tumour cell lines and primary immune cells." Biochem Pharmacol 74(11): 1587-1595.

Hirata, H., Y. Hinoda, K. Nakajima, K. Kawamoto, N. Kikuno, K. Ueno, S. Yamamura, M. S. Zaman, G. Khatri, Y. Chen, S. Saini, S. Majid, G. Deng, N. Ishii and R. Dahiya (2011). "Wnt antagonist DKK1 acts as a tumor suppressor gene that induces apoptosis and inhibits proliferation in human renal cell carcinoma." Int J Cancer 128(8): 1793-1803.

Hong, S. J., T. M. Dawson and V. L. Dawson (2004). "Nuclear and mitochondrial conversations in cell death: PARP-1 and AIF signaling." Trends Pharmacol Sci 25(5): 259-264.

Horoszewicz, J. S., S. S. Leong, T. M. Chu, Z. L. Wajsman, M. Friedman, L. Papsidero, U. Kim, L. S. Chai, S. Kakati, S. K. Arya and A. A. Sandberg (1980). "The LNCaP cell line--a new model for studies on human prostatic carcinoma." Prog Clin Biol Res 37: 115-132.

Horoszewicz, J. S., S. S. Leong, E. Kawinski, J. P. Karr, H. Rosenthal, T. M. Chu, E. A. Mirand and G. P. Murphy (1983). "LNCaP model of human prostatic carcinoma." Cancer Res 43(4): 1809-1818.

Horvath, L. G., S. M. Henshall, J. G. Kench, D. N. Saunders, C. S. Lee, D. Golovsky, P. C. Brenner, G. F. O'Neill, R. Kooner, P. D. Stricker, J. J. Grygiel and R. L. Sutherland (2004). "Membranous expression of secreted frizzled-related protein 4 predicts for good prognosis in localized prostate cancer and inhibits PC3 cellular proliferation in vitro." Clin Cancer Res 10(2): 615-625.

Page 270: Investigating the effects of, the isoflavone, Phenoxodiol

250

Horvath, L. G., J. E. Lelliott, J. G. Kench, C. S. Lee, E. D. Williams, D. N. Saunders, J. J. Grygiel, R. L. Sutherland and S. M. Henshall (2007). "Secreted frizzled-related protein 4 inhibits proliferation and metastatic potential in prostate cancer." Prostate 67(10): 1081-1090.

Hotte, S. J. and F. Saad (2010). "Current management of castrate-resistant prostate cancer." Curr Oncol 17 Suppl 2: S72-79.

Howes, J. B., P. L. de Souza, L. West, L. J. Huang and L. G. Howes (2011). "Pharmacokinetics of phenoxodiol, a novel isoflavone, following intravenous administration to patients with advanced cancer." BMC Clin Pharmacol 11: 1.

Hsing, A. W., L. Tsao and S. S. Devesa (2000). "International trends and patterns of prostate cancer incidence and mortality." Int J Cancer 85(1): 60-67.

Hu, Y., G. Cherton-Horvat, V. Dragowska, S. Baird, R. G. Korneluk, J. P. Durkin, L. D. Mayer and E. C. LaCasse (2003). "Antisense oligonucleotides targeting XIAP induce apoptosis and enhance chemotherapeutic activity against human lung cancer cells in vitro and in vivo." Clin Cancer Res 9(7): 2826-2836.

Huo, J. X., S. A. Metz and G. D. Li (2004). "p53-independent induction of p21(waf1/cip1) contributes to the activation of caspases in GTP-depletion-induced apoptosis of insulin-secreting cells." Cell Death Differ 11(1): 99-109.

Isaacs, J. T. and W. B. Isaacs (2004). "Androgen receptor outwits prostate cancer drugs." Nat Med 10(1): 26-27.

Janicke, R. U., P. Ng, M. L. Sprengart and A. G. Porter (1998). "Caspase-3 is required for alpha-fodrin cleavage but dispensable for cleavage of other death substrates in apoptosis." J Biol Chem 273(25): 15540-15545.

Janicke, R. U., M. L. Sprengart, M. R. Wati and A. G. Porter (1998). "Caspase-3 is required for DNA fragmentation and morphological changes associated with apoptosis." J Biol Chem 273(16): 9357-9360.

Jian, L. (2009). "Soy, isoflavones, and prostate cancer." Mol Nutr Food Res 53(2): 217-226.

Joannou, G. E., G. E. Kelly, A. Y. Reeder, M. Waring and C. Nelson (1995). "A urinary profile study of dietary phytoestrogens. The identification and mode of metabolism of new isoflavonoids." J Steroid Biochem Mol Biol 54(3-4): 167-184.

Johnstone, R. W., A. A. Ruefli and S. W. Lowe (2002). "Apoptosis: a link between cancer genetics and chemotherapy." Cell 108(2): 153-164.

Juhaszova, M., D. B. Zorov, S. H. Kim, S. Pepe, Q. Fu, K. W. Fishbein, B. D. Ziman, S. Wang, K. Ytrehus, C. L. Antos, E. N. Olson and S. J. Sollott (2004). "Glycogen synthase kinase-3beta mediates convergence of protection signaling to inhibit the mitochondrial permeability transition pore." J Clin Invest 113(11): 1535-1549.

Kaighn, M. E., K. S. Narayan, Y. Ohnuki, J. F. Lechner and L. W. Jones (1979). "Establishment and characterization of a human prostatic carcinoma cell line (PC-3)." Invest Urol 17(1): 16-23.

Kamsteeg, M., T. Rutherford, E. Sapi, B. Hanczaruk, S. Shahabi, M. Flick, D. Brown and G. Mor (2003). "Phenoxodiol--an isoflavone analog--induces apoptosis in chemoresistant ovarian cancer cells." Oncogene 22(17): 2611-2620.

Page 271: Investigating the effects of, the isoflavone, Phenoxodiol

251

Kantoff, P. W., S. Halabi, M. Conaway, J. Picus, J. Kirshner, V. Hars, D. Trump, E. P. Winer and N. J. Vogelzang (1999). "Hydrocortisone with or without mitoxantrone in men with hormone-refractory prostate cancer: results of the cancer and leukemia group B 9182 study." J Clin Oncol 17(8): 2506-2513.

Kawano, Y., S. Diez, P. Uysal-Onganer, R. S. Darrington, J. Waxman and R. M. Kypta (2009). "Secreted Frizzled-related protein-1 is a negative regulator of androgen receptor activity in prostate cancer." Br J Cancer 100(7): 1165-1174.

Kelly, G. E. and A. J. Husband (2003). "Flavonoid compounds in the prevention and treatment of prostate cancer." Methods Mol Med 81: 377-394.

Kerr, J. F., A. H. Wyllie and A. R. Currie (1972). "Apoptosis: a basic biological phenomenon with wide-ranging implications in tissue kinetics." Br J Cancer 26(4): 239-257.

Kessler, B. and P. Albertsen (2003). "The natural history of prostate cancer." Urol Clin North Am 30(2): 219-226.

King, T. J., L. H. Fukushima, T. A. Donlon, A. D. Hieber, K. A. Shimabukuro and J. S. Bertram (2000). "Correlation between growth control, neoplastic potential and endogenous connexin43 expression in HeLa cell lines: implications for tumor progression." Carcinogenesis 21(2): 311-315.

Kleeberger, W., G. S. Bova, M. E. Nielsen, M. Herawi, A. Y. Chuang, J. I. Epstein and D. M. Berman (2007). "Roles for the stem cell associated intermediate filament Nestin in prostate cancer migration and metastasis." Cancer Res 67(19): 9199-9206.

Klein, R., D. Brown and A. M. Turnley (2007). "Phenoxodiol protects against Cisplatin induced neurite toxicity in a PC-12 cell model." BMC Neurosci 8: 61.

Kluger, H. M., M. M. McCarthy, A. B. Alvero, M. Sznol, S. Ariyan, R. L. Camp, D. L. Rimm and G. Mor (2007). "The X-linked inhibitor of apoptosis protein (XIAP) is up-regulated in metastatic melanoma, and XIAP cleavage by Phenoxodiol is associated with Carboplatin sensitization." J Transl Med 5: 6.

Koda, M., M. Sulkowska, L. Kanczuga-Koda, J. Tomaszewski, W. Kucharczuk, T. Lesniewicz, S. Cymek and S. Sulkowski (2007). "The effect of chemotherapy on Ki-67, Bcl-2 and Bak expression in primary tumors and lymph node metastases of breast cancer." Oncol Rep 18(1): 113-119.

Korsmeyer, S. J. (1992). "Bcl-2 initiates a new category of oncogenes: regulators of cell death." Blood 80(4): 879-886.

Korsmeyer, S. J., M. C. Wei, M. Saito, S. Weiler, K. J. Oh and P. H. Schlesinger (2000). "Pro-apoptotic cascade activates BID, which oligomerizes BAK or BAX into pores that result in the release of cytochrome c." Cell Death Differ 7(12): 1166-1173.

Krajewska, M., S. Krajewski, S. Banares, X. Huang, B. Turner, L. Bubendorf, O. P. Kallioniemi, A. Shabaik, A. Vitiello, D. Peehl, G. J. Gao and J. C. Reed (2003). "Elevated expression of inhibitor of apoptosis proteins in prostate cancer." Clin Cancer Res 9(13): 4914-4925.

Kuruvilla, S., C. W. Qualls, Jr., R. D. Tyler, S. M. Witherspoon, G. R. Benavides, L. W. Yoon, K. Dold, R. H. Brown, S. Sangiah and K. T. Morgan (2003). "Effects of minimally toxic levels of carbonyl cyanide P-(trifluoromethoxy) phenylhydrazone (FCCP), elucidated through

Page 272: Investigating the effects of, the isoflavone, Phenoxodiol

252

differential gene expression with biochemical and morphological correlations." Toxicol Sci 73(2): 348-361.

Ladha, M. H., K. Y. Lee, T. M. Upton, M. F. Reed and M. E. Ewen (1998). "Regulation of exit from quiescence by p27 and cyclin D1-CDK4." Mol Cell Biol 18(11): 6605-6615.

Langley, R. R. and I. J. Fidler (2011). "The seed and soil hypothesis revisited--the role of tumor-stroma interactions in metastasis to different organs." Int J Cancer 128(11): 2527-2535.

Lee, D. H., M. Szczepanski and Y. J. Lee (2008). "Role of Bax in quercetin-induced apoptosis in human prostate cancer cells." Biochem Pharmacol 75(12): 2345-2355.

Levine, A. J. (1997). "p53, the cellular gatekeeper for growth and division." Cell 88(3): 323-331.

Levkau, B., K. J. Garton, N. Ferri, K. Kloke, J. R. Nofer, H. A. Baba, E. W. Raines and G. Breithardt (2001). "xIAP induces cell-cycle arrest and activates nuclear factor-kappaB : new survival pathways disabled by caspase-mediated cleavage during apoptosis of human endothelial cells." Circ Res 88(3): 282-290.

Li, H., X. Chen, T. Calhoun-Davis, K. Claypool and D. G. Tang (2008). "PC3 human prostate carcinoma cell holoclones contain self-renewing tumor-initiating cells." Cancer Res 68(6): 1820-1825.

Li, J., M. Xing, M. Zhu, X. Wang, M. Wang, S. Zhou, N. Li, R. Wu and M. Zhou (2008). "Glycogen synthase kinase 3beta induces apoptosis in cancer cells through increase of survivin nuclear localization." Cancer Lett 272(1): 91-101.

Li, K., Y. Li, J. M. Shelton, J. A. Richardson, E. Spencer, Z. J. Chen, X. Wang and R. S. Williams (2000). "Cytochrome c deficiency causes embryonic lethality and attenuates stress-induced apoptosis." Cell 101(4): 389-399.

Li, Y. and F. H. Sarkar (2002). "Gene expression profiles of genistein-treated PC3 prostate cancer cells." J Nutr 132(12): 3623-3631.

Li, Y., Z. Wang, D. Kong, R. Li, S. H. Sarkar and F. H. Sarkar (2008). "Regulation of Akt/FOXO3a/GSK-3beta/AR signaling network by isoflavone in prostate cancer cells." J Biol Chem 283(41): 27707-27716.

Liang, H., Q. Chen, A. H. Coles, S. J. Anderson, G. Pihan, A. Bradley, R. Gerstein, R. Jurecic and S. N. Jones (2003). "Wnt5a inhibits B cell proliferation and functions as a tumor suppressor in hematopoietic tissue." Cancer Cell 4(5): 349-360.

Liss, M. A., M. Schlicht, A. Kahler, R. Fitzgerald, T. Thomassi, A. Degueme, M. Hessner and M. W. Datta (2010). "Characterization of soy-based changes in Wnt-frizzled signaling in prostate cancer." Cancer Genomics Proteomics 7(5): 245-252.

Litvinov, I. V., A. M. De Marzo and J. T. Isaacs (2003). "Is the Achilles' heel for prostate cancer therapy a gain of function in androgen receptor signaling?" J Clin Endocrinol Metab 88(7): 2972-2982.

MacFarlane, M., W. Merrison, S. B. Bratton and G. M. Cohen (2002). "Proteasome-mediated degradation of Smac during apoptosis: XIAP promotes Smac ubiquitination in vitro." J Biol Chem 277(39): 36611-36616.

Page 273: Investigating the effects of, the isoflavone, Phenoxodiol

253

MacLachlan, T. K., N. Sang and A. Giordano (1995). "Cyclins, cyclin-dependent kinases and cdk inhibitors: implications in cell cycle control and cancer." Crit Rev Eukaryot Gene Expr 5(2): 127-156.

Madani, S. H., S. Ameli, S. Khazaei, M. Kanani and B. Izadi (2011). "Frequency of Ki-67 (MIB-1) and P53 expressions among patients with prostate cancer." Indian J Pathol Microbiol 54(4): 688-691.

Maguire, T., P. Harrison, O. Hyink, J. Kalmakoff and V. K. Ward (2000). "The inhibitors of apoptosis of Epiphyas postvittana nucleopolyhedrovirus." J Gen Virol 81(Pt 11): 2803-2811.

Mahoney, S., Berry, C., Hisheh, S., Dharmarajan, A. (2007). "X-linked Inhibitor of Apoptosis Protein Expression and Cisplatin Resistance in Ovarian Cancer." International Journal of Integrative Biology 1(1): 8.

Marieb, E., . Hoehn, K. (2012). Human Anatomy and Physiology. Sydney, Pearson.

Martin, S. J., C. P. Reutelingsperger, A. J. McGahon, J. A. Rader, R. C. van Schie, D. M. LaFace and D. R. Green (1995). "Early redistribution of plasma membrane phosphatidylserine is a general feature of apoptosis regardless of the initiating stimulus: inhibition by overexpression of Bcl-2 and Abl." J Exp Med 182(5): 1545-1556.

Mateyak, M. K., A. J. Obaya and J. M. Sedivy (1999). "c-Myc regulates cyclin D-Cdk4 and -Cdk6 activity but affects cell cycle progression at multiple independent points." Mol Cell Biol 19(7): 4672-4683.

McCarthy, P. J., S. F. Sweetman, P. G. McKenna and V. J. McKelvey-Martin (1997). "Evaluation of manual and image analysis quantification of DNA damage in the alkaline comet assay." Mutagenesis 12(4): 209-214.

McNeal, J. E., E. A. Redwine, F. S. Freiha and T. A. Stamey (1988). "Zonal distribution of prostatic adenocarcinoma. Correlation with histologic pattern and direction of spread." Am J Surg Pathol 12(12): 897-906.

McPherson, R. A., P. T. Galettis and P. L. de Souza (2009). "Enhancement of the activity of phenoxodiol by cisplatin in prostate cancer cells." Br J Cancer 100(4): 649-655.

Menezes, M. E., D. J. Devine, L. A. Shevde and R. S. Samant (2012). "Dickkopf1: a tumor suppressor or metastasis promoter?" Int J Cancer 130(7): 1477-1483.

Michaud, L. B., V. Valero and G. Hortobagyi (2000). "Risks and benefits of taxanes in breast and ovarian cancer." Drug Saf 23(5): 401-428.

Middleton, E., Jr., C. Kandaswami and T. C. Theoharides (2000). "The effects of plant flavonoids on mammalian cells: implications for inflammation, heart disease, and cancer." Pharmacol Rev 52(4): 673-751.

Minelli, A., I. Bellezza, A. Tucci, M. G. Rambotti, C. Conte and Z. Culig (2009). "Differential involvement of reactive oxygen species and nucleoside transporters in cytotoxicity induced by two adenosine analogues in human prostate cancer cells." Prostate 69(5): 538-547.

Page 274: Investigating the effects of, the isoflavone, Phenoxodiol

254

Mitchell, J. H., S. J. Duthie and A. R. Collins (2000). "Effects of phytoestrogens on growth and DNA integrity in human prostate tumor cell lines: PC-3 and LNCaP." Nutr Cancer 38(2): 223-228.

Miyamoto, H., S. Altuwaijri, Y. Cai, E. M. Messing and C. Chang (2005). "Inhibition of the Akt, cyclooxygenase-2, and matrix metalloproteinase-9 pathways in combination with androgen deprivation therapy: potential therapeutic approaches for prostate cancer." Mol Carcinog 44(1): 1-10.

Miyamoto, H., E. M. Messing and C. Chang (2004). "Androgen deprivation therapy for prostate cancer: current status and future prospects." Prostate 61(4): 332-353.

Montero, A., F. Fossella, G. Hortobagyi and V. Valero (2005). "Docetaxel for treatment of solid tumours: a systematic review of clinical data." Lancet Oncol 6(4): 229-239.

Moore, K., A. Agur, A. Dalley, V. Oxorn and M. Moore (2007). Essential Clinical Anatomy. Baltimore, Philidelphia, Lippincott Williams & Wilkins.

Mor, G., H. H. Fu and A. B. Alvero (2006). "Phenoxodiol, a novel approach for the treatment of ovarian cancer." Curr Opin Investig Drugs 7(6): 542-548.

Morre, D. J., P. J. Chueh, K. Yagiz, A. Balicki, C. Kim and D. M. Morre (2007). "ECTO-NOX target for the anticancer isoflavene phenoxodiol." Oncol Res 16(7): 299-312.

Morre, D. J., N. McClain, L. Y. Wu, G. Kelly and D. M. Morre (2009). "Phenoxodiol treatment alters the subsequent response of ENOX2 (tNOX) and growth of hela cells to paclitaxel and cisplatin." Mol Biotechnol 42(1): 100-109.

Muley, A., S. Majumder, G. K. Kolluru, S. Parkinson, H. Viola, L. Hool, F. Arfuso, R. Ganss, A. Dharmarajan and S. Chatterjee (2010). "Secreted frizzled-related protein 4: an angiogenesis inhibitor." Am J Pathol 176(3): 1505-1516.

Murphy, M. P. (2009). "How mitochondria produce reactive oxygen species." Biochem J 417(1): 1-13.

Nakagawa, T., H. Zhu, N. Morishima, E. Li, J. Xu, B. A. Yankner and J. Yuan (2000). "Caspase-12 mediates endoplasmic-reticulum-specific apoptosis and cytotoxicity by amyloid-beta." Nature 403(6765): 98-103.

Niculescu, A. B., 3rd, X. Chen, M. Smeets, L. Hengst, C. Prives and S. I. Reed (1998). "Effects of p21(Cip1/Waf1) at both the G1/S and the G2/M cell cycle transitions: pRb is a critical determinant in blocking DNA replication and in preventing endoreduplication." Mol Cell Biol 18(1): 629-643.

Pardee, A. B. (1974). "A restriction point for control of normal animal cell proliferation." Proc Natl Acad Sci U S A 71(4): 1286-1290.

Parente, P., F. Parnis and H. Gurney (2012). "Emerging and second line therapies for the management of metastatic castration-resistant prostate cancer: The Australian perspective." Asia Pac J Clin Oncol 8(1): 31-42.

Park, Y. J., K. H. Bae, B. S. Jin, H. J. Jung and J. S. Park (2012). "Is increased prostatic urethral angle related to lower urinary tract symptoms in males with benign prostatic hyperplasia/lower urinary tract symptoms?" Korean J Urol 53(6): 410-413.

Page 275: Investigating the effects of, the isoflavone, Phenoxodiol

255

Parton, M., S. Krajewski, I. Smith, M. Krajewska, C. Archer, M. Naito, R. Ahern, J. Reed and M. Dowsett (2002). "Coordinate expression of apoptosis-associated proteins in human breast cancer before and during chemotherapy." Clin Cancer Res 8(7): 2100-2108.

Peng, Y., C. X. Li, F. Chen, Z. Wang, M. Ligr, J. Melamed, J. Wei, W. Gerald, M. Pagano, M. J. Garabedian and P. Lee (2008). "Stimulation of prostate cancer cellular proliferation and invasion by the androgen receptor co-activator ARA70." Am J Pathol 172(1): 225-235.

Petronilli, V., D. Penzo, L. Scorrano, P. Bernardi and F. Di Lisa (2001). "The mitochondrial permeability transition, release of cytochrome c and cell death. Correlation with the duration of pore openings in situ." J Biol Chem 276(15): 12030-12034.

Petrylak, D. P., C. M. Tangen, M. H. Hussain, P. N. Lara, Jr., J. A. Jones, M. E. Taplin, P. A. Burch, D. Berry, C. Moinpour, M. Kohli, M. C. Benson, E. J. Small, D. Raghavan and E. D. Crawford (2004). "Docetaxel and estramustine compared with mitoxantrone and prednisone for advanced refractory prostate cancer." N Engl J Med 351(15): 1513-1520.

Puthalakath, H. and A. Strasser (2002). "Keeping killers on a tight leash: transcriptional and post-translational control of the pro-apoptotic activity of BH3-only proteins." Cell Death Differ 9(5): 505-512.

Reed, J. C. (2000). "Mechanisms of apoptosis." Am J Pathol 157(5): 1415-1430.

Reya, T., S. J. Morrison, M. F. Clarke and I. L. Weissman (2001). "Stem cells, cancer, and cancer stem cells." Nature 414(6859): 105-111.

Riedl, S. J. and G. S. Salvesen (2007). "The apoptosome: signalling platform of cell death." Nat Rev Mol Cell Biol 8(5): 405-413.

Ririe, K. M., R. P. Rasmussen and C. T. Wittwer (1997). "Product differentiation by analysis of DNA melting curves during the polymerase chain reaction." Anal Biochem 245(2): 154-160.

Risbridger, G. P., H. Wang, M. Frydenberg and A. Husband (2001). "The in vivo effect of red clover diet on ventral prostate growth in adult male mice." Reprod Fertil Dev 13(4): 325-329.

Roy, S., M. Kaur, C. Agarwal, M. Tecklenburg, R. A. Sclafani and R. Agarwal (2007). "p21 and p27 induction by silibinin is essential for its cell cycle arrest effect in prostate carcinoma cells." Mol Cancer Ther 6(10): 2696-2707.

Roy, S., R. P. Singh, C. Agarwal, S. Siriwardana, R. Sclafani and R. Agarwal (2008). "Downregulation of both p21/Cip1 and p27/Kip1 produces a more aggressive prostate cancer phenotype." Cell Cycle 7(12): 1828-1835.

Ruth, A. C. and I. B. Roninson (2000). "Effects of the multidrug transporter P-glycoprotein on cellular responses to ionizing radiation." Cancer Res 60(10): 2576-2578.

Ryan, K. M., A. C. Phillips and K. H. Vousden (2001). "Regulation and function of the p53 tumor suppressor protein." Curr Opin Cell Biol 13(3): 332-337.

Page 276: Investigating the effects of, the isoflavone, Phenoxodiol

256

Saif, M. W., E. Tytler, F. Lansigan, D. M. Brown and A. J. Husband (2009). "Flavonoids, phenoxodiol, and a novel agent, triphendiol, for the treatment of pancreaticobiliary cancers." Expert Opin Investig Drugs 18(4): 469-479.

Saladin, K. (2007). Anatomy and Physiology: The Unity of Form and Function. New York, McGraw Hill.

Salvioli, S., A. Ardizzoni, C. Franceschi and A. Cossarizza (1997). "JC-1, but not DiOC6(3) or rhodamine 123, is a reliable fluorescent probe to assess delta psi changes in intact cells: implications for studies on mitochondrial functionality during apoptosis." FEBS Lett 411(1): 77-82.

Sanna, M. G., J. da Silva Correia, Y. Luo, B. Chuang, L. M. Paulson, B. Nguyen, Q. L. Deveraux and R. J. Ulevitch (2002). "ILPIP, a novel anti-apoptotic protein that enhances XIAP-mediated activation of JNK1 and protection against apoptosis." J Biol Chem 277(34): 30454-30462.

Sapi, E., A. B. Alvero, W. Chen, D. O'Malley, X. Y. Hao, B. Dwipoyono, M. Garg, M. Kamsteeg, T. Rutherford and G. Mor (2004). "Resistance of ovarian carcinoma cells to docetaxel is XIAP dependent and reversible by phenoxodiol." Oncol Res 14(11-12): 567-578.

Scherr, D. S., E. D. Vaughan, Jr., J. Wei, M. Chung, D. Felsen, R. Allbright and B. S. Knudsen (1999). "BCL-2 and p53 expression in clinically localized prostate cancer predicts response to external beam radiotherapy." J Urol 162(1): 12-16; discussion 16-17.

Schmitt, E., A. Steyaert, G. Cimoli and R. Bertrand (1998). "Bax-alpha promotes apoptosis induced by cancer chemotherapy and accelerates the activation of caspase 3-like cysteine proteases in p53 double mutant B lymphoma Namalwa cells." Cell Death Differ 5(6): 506-516.

Schwartzman, R. A. and J. A. Cidlowski (1993). "Apoptosis: the biochemistry and molecular biology of programmed cell death." Endocr Rev 14(2): 133-151.

Senderowicz, A. M. (2004). "Targeting cell cycle and apoptosis for the treatment of human malignancies." Curr Opin Cell Biol 16(6): 670-678.

Seo, Y. J., B. S. Kim, S. Y. Chun, Y. K. Park, K. S. Kang and T. G. Kwon (2011). "Apoptotic effects of genistein, biochanin-A and apigenin on LNCaP and PC-3 cells by p21 through transcriptional inhibition of polo-like kinase-1." J Korean Med Sci 26(11): 1489-1494.

Silasi, D. A., A. B. Alvero, T. J. Rutherford, D. Brown and G. Mor (2009). "Phenoxodiol: pharmacology and clinical experience in cancer monotherapy and in combination with chemotherapeutic drugs." Expert Opin Pharmacother 10(6): 1059-1067.

Silke, J., P. G. Ekert, C. L. Day, C. J. Hawkins, M. Baca, J. Chew, M. Pakusch, A. M. Verhagen and D. L. Vaux (2001). "Direct inhibition of caspase 3 is dispensable for the anti-apoptotic activity of XIAP." Embo J 20(12): 3114-3123.

Slee, E. A., C. Adrain and S. J. Martin (1999). "Serial killers: ordering caspase activation events in apoptosis." Cell Death Differ 6(11): 1067-1074.

Slee, E. A., M. T. Harte, R. M. Kluck, B. B. Wolf, C. A. Casiano, D. D. Newmeyer, H. G. Wang, J. C. Reed, D. W. Nicholson, E. S. Alnemri, D. R. Green and S. J. Martin (1999). "Ordering the cytochrome c-initiated caspase cascade: hierarchical activation of caspases-2, -3, -6, -7, -8, and -10 in a caspase-9-dependent manner." J Cell Biol 144(2): 281-292.

Page 277: Investigating the effects of, the isoflavone, Phenoxodiol

257

So, A. I., A. Hurtado-Coll and M. E. Gleave (2003). "Androgens and prostate cancer." World J Urol 21(5): 325-337.

Srinivasula, S. M., R. Hegde, A. Saleh, P. Datta, E. Shiozaki, J. Chai, R. A. Lee, P. D. Robbins, T. Fernandes-Alnemri, Y. Shi and E. S. Alnemri (2001). "A conserved XIAP-interaction motif in caspase-9 and Smac/DIABLO regulates caspase activity and apoptosis." Nature 410(6824): 112-116.

Srivastava, R. K., Q. Chen, I. Siddiqui, K. Sarva and S. Shankar (2007). "Linkage of curcumin-induced cell cycle arrest and apoptosis by cyclin-dependent kinase inhibitor p21(/WAF1/CIP1)." Cell Cycle 6(23): 2953-2961.

Stevens, A., . Lowe, J,. Scott, I,. (2009). Core Pathology, Mosby Elsevier.

Stone, K. R., D. D. Mickey, H. Wunderli, G. H. Mickey and D. F. Paulson (1978). "Isolation of a human prostate carcinoma cell line (DU 145)." Int J Cancer 21(3): 274-281.

Straszewski-Chavez, S. L., V. M. Abrahams, E. F. Funai and G. Mor (2004). "X-linked inhibitor of apoptosis (XIAP) confers human trophoblast cell resistance to Fas-mediated apoptosis." Mol Hum Reprod 10(1): 33-41.

Sun, A., J. Tang, Y. Hong, J. Song, P. F. Terranova, J. B. Thrasher, S. Svojanovsky, H. G. Wang and B. Li (2008). "Androgen receptor-dependent regulation of Bcl-xL expression: Implication in prostate cancer progression." Prostate 68(4): 453-461.

Susin, S. A., E. Daugas, L. Ravagnan, K. Samejima, N. Zamzami, M. Loeffler, P. Costantini, K. F. Ferri, T. Irinopoulou, M. C. Prevost, G. Brothers, T. W. Mak, J. Penninger, W. C. Earnshaw and G. Kroemer (2000). "Two distinct pathways leading to nuclear apoptosis." J Exp Med 192(4): 571-580.

Susin, S. A., H. K. Lorenzo, N. Zamzami, I. Marzo, B. E. Snow, G. M. Brothers, J. Mangion, E. Jacotot, P. Costantini, M. Loeffler, N. Larochette, D. R. Goodlett, R. Aebersold, D. P. Siderovski, J. M. Penninger and G. Kroemer (1999). "Molecular characterization of mitochondrial apoptosis-inducing factor." Nature 397(6718): 441-446.

Suzuki, Y., Y. Nakabayashi, K. Nakata, J. C. Reed and R. Takahashi (2001). "X-linked inhibitor of apoptosis protein (XIAP) inhibits caspase-3 and -7 in distinct modes." J Biol Chem 276(29): 27058-27063.

Svingen, P. A., D. Loegering, J. Rodriquez, X. W. Meng, P. W. Mesner, Jr., S. Holbeck, A. Monks, S. Krajewski, D. A. Scudiero, E. A. Sausville, J. C. Reed, Y. A. Lazebnik and S. H. Kaufmann (2004). "Components of the cell death machine and drug sensitivity of the National Cancer Institute Cell Line Panel." Clin Cancer Res 10(20): 6807-6820.

Sweat, G. T. (2005). "Guiding prostate cancer treatment choices. Early detection means more options for more men." Postgrad Med 117(4): 45-50.

Tallarida, R. J. (2006). "An overview of drug combination analysis with isobolograms." J Pharmacol Exp Ther 319(1): 1-7.

Tang, Y., M. A. Khan, O. Goloubeva, D. I. Lee, D. Jelovac, A. M. Brodie and A. Hussain (2006). "Docetaxel followed by castration improves outcomes in LNCaP prostate cancer-bearing severe combined immunodeficient mice." Clin Cancer Res 12(1): 169-174.

Page 278: Investigating the effects of, the isoflavone, Phenoxodiol

258

Tannock, I. F., R. de Wit, W. R. Berry, J. Horti, A. Pluzanska, K. N. Chi, S. Oudard, C. Theodore, N. D. James, I. Turesson, M. A. Rosenthal and M. A. Eisenberger (2004). "Docetaxel plus prednisone or mitoxantrone plus prednisone for advanced prostate cancer." N Engl J Med 351(15): 1502-1512.

Tannock, I. F. and C. Lee (2001). "Evidence against apoptosis as a major mechanism for reproductive cell death following treatment of cell lines with anti-cancer drugs." Br J Cancer 84(1): 100-105.

Trosko, J. E., C. C. Chang, B. L. Upham and M. H. Tai (2004). "Ignored hallmarks of carcinogenesis: stem cells and cell-cell communication." Ann N Y Acad Sci 1028: 192-201.

Uzgare, A. R., Y. Xu and J. T. Isaacs (2004). "In vitro culturing and characteristics of transit amplifying epithelial cells from human prostate tissue." J Cell Biochem 91(1): 196-205.

Vafa, O., M. Wade, S. Kern, M. Beeche, T. K. Pandita, G. M. Hampton and G. M. Wahl (2002). "c-Myc can induce DNA damage, increase reactive oxygen species, and mitigate p53 function: a mechanism for oncogene-induced genetic instability." Mol Cell 9(5): 1031-1044.

van Es, J. H., N. Barker and H. Clevers (2003). "You Wnt some, you lose some: oncogenes in the Wnt signaling pathway." Curr Opin Genet Dev 13(1): 28-33.

van Heerde, W. L., S. Robert-Offerman, E. Dumont, L. Hofstra, P. A. Doevendans, J. F. Smits, M. J. Daemen and C. P. Reutelingsperger (2000). "Markers of apoptosis in cardiovascular tissues: focus on Annexin V." Cardiovasc Res 45(3): 549-559.

Waldman, T., C. Lengauer, K. W. Kinzler and B. Vogelstein (1996). "Uncoupling of S phase and mitosis induced by anticancer agents in cells lacking p21." Nature 381(6584): 713-716.

Wan, X., W. Ji, X. Mei, J. Zhou, J. X. Liu, C. Fang and W. Xiao (2010). "Negative feedback regulation of Wnt4 signaling by EAF1 and EAF2/U19." PLoS One 5(2): e9118.

Wan, X., J. Liu, J. F. Lu, V. Tzelepi, J. Yang, M. W. Starbuck, L. Diao, J. Wang, E. Efstathiou, E. S. Vazquez, P. Troncoso, S. N. Maity and N. M. Navone (2012). "Activation of beta-catenin signaling in androgen receptor-negative prostate cancer cells." Clin Cancer Res 18(3): 726-736.

Wang, G., J. Wang and M. D. Sadar (2008). "Crosstalk between the androgen receptor and beta-catenin in castrate-resistant prostate cancer." Cancer Res 68(23): 9918-9927.

Wang, L. G., X. M. Liu, W. Kreis and D. R. Budman (1999). "The effect of antimicrotubule agents on signal transduction pathways of apoptosis: a review." Cancer Chemother Pharmacol 44(5): 355-361.

Wang, P., S. M. Henning and D. Heber (2010). "Limitations of MTT and MTS-based assays for measurement of antiproliferative activity of green tea polyphenols." PLoS One 5(4): e10202.

Wang, R., F. Lin, X. Wang, P. Gao, K. Dong, S. H. Wei, S. Y. Cheng and H. Z. Zhang (2008). "Suppression of Bcl-xL expression by a novel tumor-specific RNA interference system inhibits proliferation and enhances radiosensitivity in prostatic carcinoma cells." Cancer Chemother Pharmacol 61(6): 943-952.

Page 279: Investigating the effects of, the isoflavone, Phenoxodiol

259

Wasylishen, A. R. and L. Z. Penn (2010). "Myc: the beauty and the beast." Genes Cancer 1(6): 532-541.

Watcharasit, P., G. N. Bijur, J. W. Zmijewski, L. Song, A. Zmijewska, X. Chen, G. V. Johnson and R. S. Jope (2002). "Direct, activating interaction between glycogen synthase kinase-3beta and p53 after DNA damage." Proc Natl Acad Sci U S A 99(12): 7951-7955.

Widyarini, S., N. Spinks, A. J. Husband and V. E. Reeve (2001). "Isoflavonoid compounds from red clover (Trifolium pratense) protect from inflammation and immune suppression induced by UV radiation." Photochem Photobiol 74(3): 465-470.

Wu, L. Y., T. De Luca, T. Watanabe, D. M. Morre and D. J. Morre (2011). "Metabolite modulation of HeLa cell response to ENOX2 inhibitors EGCG and phenoxodiol." Biochim Biophys Acta 1810(8): 784-789.

Wyllie, A. H., J. F. Kerr and A. R. Currie (1980). "Cell death: the significance of apoptosis." Int Rev Cytol 68: 251-306.

Xie, Q., T. Lin, Y. Zhang, J. Zheng and J. A. Bonanno (2005). "Molecular cloning and characterization of a human AIF-like gene with ability to induce apoptosis." J Biol Chem 280(20): 19673-19681.

Xie, W., Y. He, D. Huo, Y. Huang and W. Wu (2010). "A mimic of phosphorylated prolactin inhibits human breast cancer cell proliferation via upregulation of p21 waf1." Med Oncol 27(4): 1340-1345.

Yagiz, K., L. Y. Wu, C. P. Kuntz, D. James Morre and D. M. Morre (2007). "Mouse embryonic fibroblast cells from transgenic mice overexpressing tNOX exhibit an altered growth and drug response phenotype." J Cell Biochem 101(2): 295-306.

Yang, F., A. von Knethen and B. Brune (2000). "Modulation of nitric oxide-evoked apoptosis by the p53-downstream target p21(WAF1/CIP1)." J Leukoc Biol 68(6): 916-922.

Yin, X. Y., L. Grove, N. S. Datta, M. W. Long and E. V. Prochownik (1999). "C-myc overexpression and p53 loss cooperate to promote genomic instability." Oncogene 18(5): 1177-1184.

Yu, F., R. N. Watts, X. D. Zhang, J. M. Borrow and P. Hersey (2006). "Involvement of BH3-only proapoptotic proteins in mitochondrial-dependent Phenoxodiol-induced apoptosis of human melanoma cells." Anticancer Drugs 17(10): 1151-1161.

Yu, X., Y. Wang, D. J. DeGraff, M. L. Wills and R. J. Matusik (2011). "Wnt/beta-catenin activation promotes prostate tumor progression in a mouse model." Oncogene 30(16): 1868-1879.

Zamzami, N., C. El Hamel, C. Maisse, C. Brenner, C. Munoz-Pinedo, A. S. Belzacq, P. Costantini, H. Vieira, M. Loeffler, G. Molle and G. Kroemer (2000). "Bid acts on the permeability transition pore complex to induce apoptosis." Oncogene 19(54): 6342-6350.

Zamzami, N., D. Metivier and G. Kroemer (2000). "Quantitation of mitochondrial transmembrane potential in cells and in isolated mitochondria." Methods Enzymol 322: 208-213.

Zhang, D. Y., H. J. Wang and Y. Z. Tan (2011). "Wnt/beta-catenin signaling induces the aging of mesenchymal stem cells through the DNA damage response and the p53/p21 pathway." PLoS One 6(6): e21397.

Page 280: Investigating the effects of, the isoflavone, Phenoxodiol

260

Zhao, L., M. G. Wientjes and J. L. Au (2004). "Evaluation of combination chemotherapy: integration of nonlinear regression, curve shift, isobologram, and combination index analyses." Clin Cancer Res 10(23): 7994-8004.

Zou, H., R. Yang, J. Hao, J. Wang, C. Sun, S. W. Fesik, J. C. Wu, K. J. Tomaselli and R. C. Armstrong (2003). "Regulation of the Apaf-1/caspase-9 apoptosome by caspase-3 and XIAP." J Biol Chem 278(10): 8091-8098.

Page 281: Investigating the effects of, the isoflavone, Phenoxodiol

261

1 0 . A P P END ICES