the impact of veph1 on ovarian cancer tumour …...thomasina spybey, master of science graduate...

114
THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR PROGRESSION IN A XENOGRAFT MOUSE MODEL by THOMASINA SPYBEY A thesis submitted in conformity with the requirements for the degree of Master of Science Graduate Department of Physiology University of Toronto © Copyright by Thomasina Spybey 2013

Upload: others

Post on 29-Sep-2020

6 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR

PROGRESSION IN A XENOGRAFT MOUSE MODEL

by

THOMASINA SPYBEY

A thesis submitted in conformity with the requirements

for the degree of Master of Science

Graduate Department of Physiology

University of Toronto

© Copyright by Thomasina Spybey 2013

Page 2: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

ii

THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR

PROGRESSION IN A XENOGRAFT MOUSE MODEL

Thomasina Spybey, Master of Science

Graduate Department of Physiology, University of Toronto

Toronto, Ontario, Canada

2013

ABSTRACT

Epithelial ovarian cancer has the greatest death to incidence ratio of all gynecological cancers in

Canada. The FoxO, TGF-β, and Hippo signalling pathways are implicated in ovarian

tumourigenesis; ventricular zone-expressed PH domain protein (Veph1) may be a novel

modulator of these pathways. This thesis addresses the objectives of determining the

physiological murine tissue expression of Veph1A/B isoforms, and establishing the effect of

Veph1 expression on ovarian tumourigenesis. Veph1A and Veph1B transcripts were identified in

whole embryos; whereas, only Veph1A transcripts were present in adult tissues. Veph1

expression significantly reduced tumour growth rate in a xenograft mouse model. This was

associated with reduced microvessel density and increased necrotic tumour area with no effect

on proliferation index. This study suggests that Veph1 may be involved in angiogenesis. Future

investigation regarding the effect of Veph1 expression on metastases and ascites formation will

be necessary to establish its role in ovarian tumourigenesis and survival.

Page 3: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

iii

ACKNOWLEDGEMENTS

The completion of this program and thesis would not have been possible without the help

of so many people along the way.

I would first like to express my sincere appreciation for the guidance and support my

advisor Dr. Ted Brown has provided me with throughout these two years. His dedication and

genuine interest in science is admirable. I am grateful to have had a well-rounded supervisor who

is both intelligent and compassionate. Dr. Brown, I am very appreciative to have been a student

in your lab; thank you.

I would next like to thank Dr. Alexandra Kollara. I truly believe that I would not have

been able to successfully complete this program and thesis without her support. She has seen me

express happiness, sadness, annoyance (yeast), joy, success, and every emotion in between

throughout my time in the lab. Alex, I am very grateful for the contributions you have made to

this project and for being my lab mum.

Thank you to Drs. Harry Elsholtz, Norman Rosenblum, and Maurice Ringuette of my

committee for the collective advice they have contributed throughout this program. Further, I

would like to thank Dr. Elsholtz for his involvement in my final defense, along with Drs. Denise

Belsham, Ashley Bruce, and Ian Rogers.

I would also like to thank Dr. Blaise Clarke for his assistance with data analyses and

contributions to this thesis. The many hours we spent counting blood vessels was a daunting task

and I appreciate his patience and enthusiasm.

I would like to thank the past and present students of the Brown Lab including Prem,

Angela, Crystal, and Soyeon. I am grateful for the conversations, support, and memories we have

shared. A special thank you goes to Prem as she has helped me tremendously with the many

questions I have asked her on a daily basis. I am appreciative of the new friendships I have

formed with several past and present students on the 6th

floor, including Tania, Abhi, Bojana,

Taline, Sarah, Nicole, and all others. Knowing that we were all tackling the same challenge was

very comforting and the memories I have with each of you are unforgettable.

Page 4: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

iv

I would like to further thank Dragana Vukasovic for patiently welcoming me to

constantly ask her for room availability; the staff at the Centre for Cellular and Biomolecular

Research Animal Facility for their training and advice throughout the mouse project; and the

staff at Pathology Services of the Toronto Centre for Phenogenomics for their valuable advice

and answers to my many questions.

I would like to express a special thank you to Anna, Tania, and Gabriel. You have each

been instrumental in maintaining my sanity throughout this program and I will always appreciate

your support. Further I would like to thank my second family (Cathie, Bailey, Shannon, and

Jarred) for believing in me always. I would also like to thank my mum and Craig for their

support and for listening to me complain and rejoice for hours on Skype about science.

Lastly, I want to acknowledge the mice that were involved in my project that have

contributed towards developing our knowledge regarding Veph1. I do not know what the future

of this protein holds, but I am grateful to be a part of uncovering its role and hopeful clinical

application.

Page 5: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

v

2.1. DEDIC

I dedicate this thesis to my beautiful mum. Mum, you are very special to me and I

appreciate the strength and unconditional love you have always provided me with.

DEDICATION

Page 6: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

vi

TABLE OF CONTENTS

ABSTRACT ................................................................................................................................... ii

ACKNOWLEDGEMENTS ........................................................................................................ iii

DEDICATION................................................................................................................................v

TABLE OF CONTENTS ............................................................................................................ vi

LIST OF TABLES ..................................................................................................................... viii

LIST OF FIGURES ..................................................................................................................... ix

ABBREVIATIONS ...................................................................................................................... xi

Chapter 1 LITERATURE REVIEW ...........................................................................................1

1.1 Overview .............................................................................................................................1

1.2 Epithelial Ovarian Cancer ................................................................................................2

1.2.1 Epidemiology and Stage-Dependent Survival .......................................................... 2

1.2.2 Heterogeneity of Epithelial Ovarian Cancer ............................................................. 3

1.2.3 Ovarian Tumour Progression .................................................................................... 6

1.2.4 Theory of High-Grade Serous Carcinoma Pathogenesis .......................................... 8

1.2.5 Theory of Clear Cell Carcinoma Pathogenesis ....................................................... 12

1.3 Veph1 ................................................................................................................................14

1.3.1 Discovery, Structure, Function ............................................................................... 14

1.3.2 Veph1 Involvement in FoxO Signalling ................................................................. 17

1.3.3 Veph1 Involvement in TGF-β Signalling ............................................................... 22

1.3.4 Veph1 Involvement in Hippo Signalling ................................................................ 25

1.3.5 Veph1 Expression in Epithelial Ovarian Cancer .................................................... 26

1.4 Thesis Hypothesis and Objectives ..................................................................................27

Chapter 2 MATERIALS AND METHODS ..............................................................................29

2.1 Tissue Culture ....................................................................................................................29

2.2 Tissue Specimen Collection ...............................................................................................29

2.3 Animals ..............................................................................................................................29

2.4 Tumour Model ...................................................................................................................31

2.5 RNA Extraction and Reverse Transcriptase-Polymerase Chain Reaction (RT-PCR) .......32

2.6 Protein Extraction and Western Blot Analysis ..................................................................32

2.7 Immunohistochemistry ......................................................................................................37

2.8 Double Terminal Deoxynucleotidyl Transferase dUTP Nick End Labeling (TUNEL)

and Cleaved Caspase-3 Staining ........................................................................................39

2.9 Immunohistochemistry Imaging and Quantitation ............................................................40

2.10 SOD2 Activity Assay .........................................................................................................41

2.11 Statistical Analyses ............................................................................................................42

Page 7: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

vii

Chapter 3 RESULTS ...................................................................................................................43

3.1 Veph1 Expression in Murine Embryos and Adult Tissues ..........................................43

3.1.1 Veph1A and Veph1B mRNA expression in murine whole embryos and adult

tissues……………................................................................................................................. 43

3.1.2 Veph1A protein expression and distribution in murine adult tissues ................... ..43

3.2 Veph1 Expression Effect on Ovarian Tumour Progression ........................................48

3.2.1 Veph1 effect on tumour formation ......................................................................... 48

3.2.2 Veph1 effect on tumour growth rate and survival .................................................. 62

3.2.3 Veph1 effect on proliferative index ........................................................................ 62

3.2.4 Veph1 effect on necrotic tumour area ..................................................................... 70

3.2.5 Veph1 effect on microvessel density ...................................................................... 70

3.2.6 Veph1 effect on SOD2 protein expression and activity .......................................... 75

Chapter 4 DISCUSSION .............................................................................................................82

4.1 Summary of Findings .........................................................................................................82

4.2 Widespread Physiological Murine Expression of Veph1 ..................................................82

4.3 Veph1 Impedes SKOV3 Primary Tumour Growth ...........................................................86

4.4 Future Directions ...............................................................................................................91

Chapter 5 REFERENCES ..........................................................................................................93

Page 8: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

viii

LIST OF TABLES

Table 1 - Primers used for RT-PCR analyses of Veph1A, Veph1B, and ACTB. 33

Table 2 - Antibodies used for Western blot analyses. 36

Table 3 - Antibodies used for immunohistochemical staining. 38

Table 4 - Summary of Veph1A mRNA and protein expression, and protein

localization, in murine adult male and female reproductive tissues. 56

Table 5 - Summary of Veph1A mRNA and protein expression, and protein localization

in murine adult non-reproductive tissues. 57

Page 9: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

ix

LIST OF FIGURES

Figure 1 - Schematic representation of human and murine Veph1 protein isoforms. 15

Figure 2 - Schematic illustrating proposed mechanisms by which Veph1 modulates

FoxO, TGF-β, and Hippo signalling. 19-20

Figure 3- Illustration of the pMEP5-3xFlag-Veph1 expression vector. 30

Figure 4 - Schematic illustrating RT-PCR primer location within Veph1A and Veph1B

transcripts. 34

Figure 5 - Veph1A and Veph1B transcript expression in murine whole embryos. 44

Figure 6 - Veph1A and Veph1B transcript expression in murine adult reproductive and

non-reproductive tissues. 45

Figure 7 - Homology between murine and human Veph1 protein sequences and

summary of epitopes of the human Veph1 protein targeted by commercially available

antibodies. 46-47

Figure 8 - Veph1A localization in murine adult female reproductive tissues. 49

Figure 9 - Veph1A localization in murine adult male reproductive tissues. 50-51

Figure 10 - Veph1A localization in murine adult non-reproductive tissues. 52-53

Figure 11 - Veph1A localization in murine adult non-reproductive tissues. 54-55

Figure 12 - Veph1A protein expression in murine adult male and female reproductive

tissues, as determined by Western blot analyses. 58-59

Figure 13 - Veph1A protein expression in murine adult non-reproductive tissues, as

determined by Western blot analyses. 60-61

Figure 14 - SKOV3-M and SKOV3-Ve tumour formation. 63

Figure 15 - ZnSO4-supplemented drinking water induction of Flag-Veph1. 64-65

Figure 16 - SKOV3-M and SKOV3-Ve tumour growth rate and survival. 66-67

Figure 17 - SKOV3-M and SKOV3-Ve tumour nuclear Ki-67 proliferative labeling

index. 68

Figure 18 - SKOV3-M and SKOV3-Ve tumour nuclear PCNA proliferative labeling

index. 69

Figure 19 - SKOV3-M and SKOV3-Ve tumour necrotic area and liquid accumulation. 71

Figure 20 - SKOV3-M and SKOV3-Ve tumour double TUNEL and cleaved caspase-3

staining. 72-73

Figure 21 - SKOV3-M and SKOV3-Ve tumour microvessel density as determined by

CD34 staining. 74

Figure 22 - SKOV3-M and SKOV3-Ve tumour microvessel density as determined by

CD31 staining. 76

Figure 23 - SKOV3-M and SKOV3-Ve SOD2 expression and activity in vitro. 77-78

Figure 24 - SOD2 protein expression in SKOV3-M and SKOV3-Ve tumours, as

determined by Western blot analysis. 79-80

Page 10: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

x

Figure 25 - SOD2 protein expression in SKOV3-M and SKOV3-Ve tumours, as

determined by immunohistochemical analysis. 81

Figure 26 - Survival of serous ovarian cancer patients with and without VEPH1

alterations. 90

Page 11: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

xi

ABBREVIATIONS

3-D 3-Dimensional

ACTB β-actin

AEC 3-amino-9-ethylcarbazole

Akt V-akt murine thymoma viral oncogene homolog 1

ALK Activin receptor-like kinase

ANOVA Analysis of variance

AP-1 Activator protein 1

ARID1A AT-rich interactive domain-containing protein 1A

BALB/c Bagg Albino/c

BAMBI BMP and activin membrane-bound inhibitor homolog

BCA Bicinchoninic acid

Bcl B-cell lymphoma

Bim Bcl2-like protein 11

BMP Bone morphogenetic protein

BRCA1/2 Breast cancer type 1 susceptibility protein 1/2

CA-125 Cancer antigen-125

CAFs Cancer-associated fibroblasts

CD31/34 Cluster of differentiation 31/34

CdCl2 Cadmium chloride

CDH1 E-cadherin

c-Myc Myelocytomatosis oncogene

CTGF Connective tissue growth factor

DAB 3,3-diaminobenzidine

DAB2 Disabled homolog 2

DACH1 Dachshund homolog 1

DDW Double distilled water

DICER Dicer I ribonuclease type III

ECL Enhanced chemiluminescence

EGTA Ethylene glycol tetraacetic acid

EndMT Endothelial-mesenchymal transition

EMT Epithelial-mesenchymal transition

EOC Epithelial ovarian cancer

Ets-1 v-ets erythroblastosis virus E26 oncogene homolog 1

EZH2 Zeste homolog 2

Page 12: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

xii

FBS Fetal bovine serum

FoxN1 Forkhead box N1

FoxO Forkhead box sub-group O

FTE Fallopian tube epithelium

G418 Geneticin

GAPDH Glyceraldehyde 3-phosphate dehydrogenase

GFP Green fluorescent protein

H2O2 Hydrogen peroxide

HEPES 4-2-hydroxyethyl-1-piperazineethanesulfonic acid

HGSC High-grade serous carcinoma

HIF1α Hypoxia-inducible factor 1

H-Ras v-Ha-ras Harvey rat sarcoma viral oncogene homolog

HRP Horseradish peroxidase

IHC Immunohistochemical

JNK c-Jun N-terminal kinases

Ki-67 Ki-67 antigen

KRAS V-Ki-ras2 Kirsten rat sarcoma viral oncogene homolog

LATS Large tumour suppressor

LEF1 Lymphoid enhancer-binding factor 1

MMP Matrix metalloproteinase

MST Mammalian STE-20-like kinase

MT1-MMP Membrane-type I matrix metalloproteinase

mTOR Mammalian target of rapamycin

MV Microvessels

MVD Microvessel density

NADPH Nicotinamide adenine dinucleotide phosphate-oxidase

NOS3 Nitric oxide synthase

OSE Ovarian surface epithelium

PAGE Polyacrylamide gel electrophoresis

PBS Phosphate buffered saline

PCNA Proliferative cell nuclear antigen

PH Pleckstrin homology

PI3K Phosphatidylinositide 3-kinase

PIK3CA Phosphatidylinositol-4,5-bisphosphate 3-kinase, catalytic

subunit alpha

PTEN Phosphatase and tensin homolog

Puma Bcl2 binding component 3

Page 13: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

xiii

PVDF Polyvinylidene fluoride

RB1 Retinoblastoma 1

RIPA Radioimmunoprecipitation assay

ROS Reactive oxygen species

RT-PCR Reverse transcriptase-polymerase chain reaction

SC-Goat-PC Santa Cruz goat polyclonal

SC-Rab-PC Santa Cruz rabbit polyclonal

SDS Sodium dodecyl sulfate

SEM Standard error of mean

Sigma-Rab-PC Sigma rabbit polyclonal

SMAD Mothers against decapentaplegic homolog 2/3

SNK Student-Newman-Keuls

SOD Superoxide dismutase

Ta Annealing temperature

TAZ Transcriptional coactivator with PDZ-binding motif

TβRI TGF-β receptor 1

TCF4 Transcription factor 4

TGF-β Transforming growth factor β

TGFBR3 Betaglycan

TMEPAI Transmembrane prostate androgen-induced protein

TP53 Tumour protein 53

Trp53 Transformation-related protein 53

Tsc1/2 Tuberous sclerosis 1/2

TUNEL Double terminal deoxynucleotidyl transferase dUTP nick end

labeling

TVS Transvaginal ultrasound

uPA Urokinase plasminogen activator

VEGF Vascular endothelial growth factor

Veph1 Ventricular zone expressed PH domain protein 1

Wnt Wingless-type

Wts Warts

YAP Yes-associated protein

ZnSO4 Zinc sulfate

Page 14: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

1

Chapter 1 LITERATURE REVIEW

1.1 Overview

In Canada there were approximately 2,600 new cases of epithelial ovarian cancer (EOC)

diagnosed and 1,750 ovarian cancer-related deaths reported in 2012 [1]. This disease is often

diagnosed at an advanced stage (III/IV) due to a lack of specific symptoms and biomarkers of

early stage disease. While late stage EOC has a 5-year survival rate of less than 30%, Stage I

EOC has a 5-year survival rate of greater than 90%, underscoring the need for strategies to

increase the occurrence of early diagnosis [2, 3]. An improved understanding of signalling events

underlying ovarian tumourigenesis is required to improve overall survival.

Veph1 (ventricular zone expressed PH domain protein 1) may be a novel protein

involved in ovarian tumourigenesis. A role for Veph1 in EOC is strongly implicated by a recent

genome wide analysis of copy number in 68 primary human EOCs; VEPH1 copy number and

mRNA expression were significantly increased in >40% of these samples [4]. An independent

study examining 12 human ovarian clear cell carcinoma cell lines identified an increased copy

number of VEPH1 in 7 of these cell lines [5]. Furthermore, ovarian cancer cell lines ES2, HEY,

and OVCA429, which are highly motile and exhibit invasive properties [6, 7], have high levels

of Veph1 protein, as compared to SKOV3 and OVCAR3 ovarian cancer cell lines that have non-

detectable Veph1 levels and a less invasive phenotype [6, 7]. Our lab has recently generated

microarray data indicating Veph1 expression alters genes that are partially regulated by the

Forkhead box sub-group O (FoxO), Hippo, and transforming growth factor-β (TGF-β) signalling

pathways, which have both ovarian tumour suppressing and promoting functions. This is

supported by Teleman et al. [8] who have shown that Melted, the Drosophila homolog of Veph1,

impairs FoxO activity [8]. Further studies in Drosophila indicate that Melted inhibits the

expression of tumour suppressing Hippo signalling mediator warts (wts; large tumour

suppressor 1/2 (LATS1/2) homolog) in photoreceptors [9-11]. In addition, studies in our lab

indicate Veph1 inhibits canonical TGF-β SMAD2/3-dependent (mothers against decapentaplegic

homolog 2/3; SMAD2/3) signalling (Shathasivam et al., in preparation). As Veph1 may be a

modulator of several signalling pathways that have tumour suppressing and promoting functions,

determination of the net effect of Veph1 on ovarian tumourigenesis is warranted.

Page 15: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

2

The studies presented in this thesis begin by addressing the physiological murine

expression of Veph1 mRNA and protein in whole embryos and select adult tissues. Furthermore,

I examined the effect of Veph1 on ovarian tumour formation and progression using a xenograft

mouse model. In this thesis, I present evidence that Veph1 mRNA expression in adult murine

tissues is more widespread than previously indicated in the literature [12] and that evaluation of

Veph1 protein expression is not attainable with current commercially available antibodies. I also

provide evidence that Veph1 is a novel disruptor of primary tumour growth rate and promotes

survival. This may be due to an induced reduction in microvessel density (MVD) by Veph1 that

could account for an increase in necrotic tumour area.

1.2 Epithelial Ovarian Cancer

1.2.1 Epidemiology and Stage-Dependent Survival

Ovarian cancer affects approximately 204,000 women worldwide each year and is the

fifth deadliest malignancy among women in North America [13]. In Canada, there were 2,600

new cases of ovarian cancer diagnosed and 1,750 deaths reported in 2012 [1]. Although this is a

relatively uncommon cancer, the death to incidence ratio is the highest of all gynecological

cancers in Canada [1] and the world [14]. The high mortality rate associated with this disease is

largely due to the frequency of advanced stage (stage III and IV) EOC detection: more than 80%

of women are diagnosed when metastases are present in the abdomen or beyond the peritoneal

cavity. Upon initial chemotherapeutic treatment, approximately 80% of patients with advanced

stage EOC will be responsive; however, up to 70% of patients will experience a recurrence of

disease [2, 3]. Ultimately, up to 90% of these patients will succumb to their disease as malignant

cells acquire resistance to existing chemotherapeutic options [2]. Conversely, patients diagnosed

at earlier stages (stage I and II), when malignancy is confined to the ovaries, have a 5-year

survival rate of greater than 90%; however, fewer than 20% of cases are diagnosed at this point

in disease progression [3, 15, 16]. Delayed detection is largely accountable to a lack of specific

symptoms or effective screening tests to indicate early stage disease [16, 17]. Novel screening

approaches can be developed once a thorough understanding of the initiating events in ovarian

tumourigenesis is established.

Delayed detection is due in part to the vague symptoms associated with this disease that

are poor indicators of ovarian cancer. Retrospective studies indicate that approximately 95% of

Page 16: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

3

patients who are diagnosed with ovarian cancer experience one or more symptoms prior to

diagnosis [18]. Unfortunately, these symptoms- which include bloating, gas, nausea, diarrhea,

and fatigue- are common and poor specific indicators of disease. Women will often delay

seeking medical attention or will be initially misdiagnosed with gastrointestinal issues, allowing

tumour progression to continue [19]. Researchers from the Department of Obstetrics and

Gynecology at the University of Washington are currently conducting a clinical trial using

symptom-triggered screening for ovarian cancer [18]. This model involves patients completing

questionnaires regarding known symptoms. Women who screen positive using this symptom

index undergo the current standard of care for ovarian cancer testing: cancer antigen-125 (CA-

125) serum level measuring and transvaginal ultrasound (TVS). CA-125 is a poorly sensitive and

specific marker as not all ovarian cancer patients have elevated serum levels and benign

conditions can also increase serum CA-125; however, it is the only current screening option

available until novel serum markers of early disease are developed [18]. An additional

contributor to the poor survival associated with ovarian cancer is the complex nature of the

disease.

1.2.2 Heterogeneity of Epithelial Ovarian Cancer

“Ovarian cancer” is an overall name referring to a multitude of subtypes that differ in

terms of their clinical course, genetic mutations involved, and morphological phenotype. Upon

diagnosis, the current ovarian cancer treatment consists of surgical cytoreduction (debulking) in

an effort to remove the maximal amount of malignant tissue. This is followed by adjuvant

platinum-based and taxane drug combination chemotherapy [2]. The most frequently employed

combination is intravenous cisplatin and paclitaxel with additional drugs used depending on

tumour response. All ovarian malignancies are currently treated with similar therapeutic

approaches despite significant genomic differences between ovarian tumour subtypes.

Ovarian tumours consist of three main categories- germ cell, sex cord-stromal, or

epithelial cell- according to the ovarian structures the tumours appear to originate from. Ovarian

tumours of non-epithelial origin are less common and frequently benign [20]. Germ cell tumours

are derived from primordial germ cells and account for approximately 25% of ovarian tumours

and 3-7% of malignant ovarian tumours. Sex cord-stromal tumours originate from follicular -

thecal and granulosa- and stromal cells. These tumours are the least common, comprising only

Page 17: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

4

8% of all ovarian tumours and 7% of malignant tumours. The most common category of ovarian

tumours is epithelial; these account for approximately 60% of all ovarian tumours. Importantly,

90% of epithelial tumours are malignant and thus this category is the most predominant and

lethal of all ovarian tumours [20]. Epithelial tumours are organized by surgical stage and

histological tumour subtype, which is further categorized as benign, malignant or borderline, and

low- or high-grade based on tumour differentiation [20]. Of the eight prominent EOC histotypes,

serous, endometrioid, mucinous, and clear cell are the most common [16]. These histotypes

differ by their unique genetic alterations that promote malignant transformation and underlie

their unique tumour progression [21].

The clear cell histotype comprises less than 3% of all EOCs. These tumours often present

as high-grade and commonly harbour inactivation mutations in the tumour suppressor gene

ARID1A, encoding AT-rich interactive domain-containing protein 1A (ARID1A) and oncogenic

activation of PIK3CA, encoding the catalytic p110α subunit of phosphatidylinositide 3-kinase

(PI3K) [16, 22]. Tumours of the endometrioid histotype represent approximately 7% of EOCs.

These tumours present as either high- or low-grade and often harbour activating mutations in

CTNNB1, the gene encoding β-catenin, and inactivating mutations of PTEN, the gene encoding

phosphatase and tensin homolog (PTEN). The mucinous histotype is the second most common

subtype which comprises 10% of patients with EOC. These tumours are commonly low-grade

and frequently harbour inactivating mutations in PTEN and oncogenic activation of V-Ki-ras2

Kirsten rat sarcoma viral oncogene homolog (KRAS) [16, 22].

The most common subtype, accounting for 70% of EOC cases, is serous. Serous tumours

present as either high- or low-grade. High-grade serous carcinomas (HGSC) are most common,

composing greater than 90% of the serous subtype. Over 75% of HGSC cases are diagnosed at

advanced stages when survival rates are low. The poor prognosis associated with this subtype is

largely due to the rapid proliferative and metastatic capacities of these tumours. In addition, these

tumours are genetically highly unstable with greater than 95% of tumours harbouring

inactivating tumour protein 53 (TP53) mutations and 40-50% exhibiting breast cancer type 1

susceptibility protein 1 (BRCA1) and/or BRCA2 inactivation via inherited mutation or epigenetic

inhibition [21, 23]. BRCA1/2 and p53 proteins are tumour suppressors with functions that

include: maintaining genomic stability, regulating cell cycle checkpoint control and apoptosis,

Page 18: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

5

and promoting double-strand DNA break repair [23]. Recently it was revealed that HGSC itself

may have several subtypes as well. The Cancer Genome Atlas analyzed 489 HGSC tumours for

mutations, copy number alterations, and variations in miRNA expression. As expected, 96% of

tumours harbored TP53 mutations. In addition, four HGSC subtypes were identified based on

gene expression clusters categorized as immunoreactive, differentiated, proliferative, and

mesenchymal [24]. Independent genomic analyses reveal that the HGSC and high-grade

endometrioid histotypes are composed of subtypes that affect survival [25]. The fact that EOC is

composed of several unique subtypes offers an explanation as to why the current blanket

approach to treatment is not effective.

Upon diagnosis, the current ovarian cancer treatment strategy consists of surgical

cytoreduction, in an effort to remove the maximal amount of malignant tissue, followed by

adjuvant platinum-based and taxane drug combination chemotherapy [2]. As EOC is composed

of several histotypes that differ by their unique genetic alterations, chemoresponse and prognosis

between histotypes varies [21]. Five-year survival rates associated with mucinous ovarian

tumours varies from 83% at stage I to 9% at stage IV, similar to endometrioid tumours with a

five-year survival rate of 78% at stage I and 6% at stage IV. This indicates that both mucinous

and endometrioid have similar chemotherapy response to drugs that are currently used. By

contrast, stage I chemotherapy response rate for clear cell tumours can be as low as 15%,

whereas serous tumours have an approximately 70% response rate. The limited chemoresponse

of early stage clear cell tumours accounts for the reduced five-year stage I survival rate of 60%,

as compared to serous tumours at 80% [26]. The poor stage I prognosis of clear cell patients is of

importance as approximately 48% of patients with clear cell carcinoma are diagnosed at stages I

and II [26].

HGSC is the most common ovarian cancer subtype with greater than 75% of cases

diagnosed at advanced stages when five-year survival rates are as low as 25% for stage III

patients and 9% for stage IV patients [20]. Upon initial treatment, approximately 80% of patients

with advanced stage EOC will be responsive [2]; however, up to 70% of patients will experience

a recurrence of disease due to acquired resistance, accounting for the high mortality rates

associated with all subtypes at advanced stage disease [3]. Evidently, novel therapeutic

approaches are necessary to treat advanced stages and it is likely that subtype-specific treatment

Page 19: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

6

will be of most importance due to the genomic variation between histotypes. A preferable

approach to improve overall ovarian cancer survival is to detect this disease during early tumour

progression at which point survival is dramatically improved. Development of detection

strategies requires a thorough understanding of the initiating and tumour progressing steps of

each histotype.

1.2.3 Ovarian Tumour Progression

Although there are significant genomic differences between the ovarian cancer subtypes,

tumour progression often follows a similar pattern. Tumour expansion will not exceed 1-2 mm if

angiogenesis is not sufficient to provide nutrient delivery and waste removal [27]. Tumour

progression of most cancer types begins with restricted primary tumour growth until

vascularization initiates a period of rapid proliferation, local invasion, and ultimate metastasis

due to newly formed, immature, and leaky capillaries that allow entry of tumour cells [27].

Unique to epithelial ovarian cancer is the implantation of malignant tissue onto the ovarian

surface often simultaneously with peritoneal implants. Most cancers require several crucial steps

of intra- and extra-vasation in order to metastasize to distant organs via the circulation; however,

ovarian cancer metastases rarely disseminate via the vasculature and thus are infrequently found

outside the peritoneal cavity [28]. All EOC histotypes have a similar passive mechanism of

metastasis; ovarian cancer cells detach from the primary tumour as single cells or multi-cellular

aggregates (spheroids) that are carried by the physiological movement of the peritoneal fluid to

the peritoneum and the omentum [28]. Epithelial-to-mesenchymal transition (EMT) of ovarian

cancer cells facilitates the detachment of these cells as an initiating event in metastasis.

EMT involves acquisition of mesenchymal properties and the loss of epithelial properties,

such as the reduction in E-cadherin-mediated cell-cell interactions, thereby promoting invasion.

Malignant ovarian cancer cells can detach as single cells or spheroid aggregates [28, 29]. Studies

indicate that ovarian cancer cells grown in three-dimensional (3-D) cultures exhibit enhanced

resistance to anoikis (anchorage-independent programmed cell death) when treated with

chemotherapeutic agents, providing a survival advantage of cell aggregates, as compared to

single cells [30, 31]. Metastatic cells attach to collagen I of exposed peritoneal extracellular

matrix and invade the mesothelium where tumour growth continues [28, 29, 32]. Ovarian cancer

cells rarely invade deeper than the mesothelial layers of the peritoneum [28].

Page 20: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

7

Although metastasis via the vasculature is not a common characteristic of EOCs,

angiogenesis is an imperative process that supports cancer cell survival. Angiogenesis is the

development of new vessels from pre-existing ones; vascularization of malignant tissue is critical

in order to deliver nutrients and remove waste. Pathological angiogenesis is initiated by the

secretion of growth factors and proteolytic enzymes from tumour cells into the interstitium that

act on endothelial cells and basement membranes to remodel existing vessels and stimulate the

release of endothelial progenitor stem cells from the bone marrow to form new vessels [27].

Angiogenesis involves an initial activation phase that is comprised of basement membrane

degradation, increased vascular permeability, and increased endothelial cell proliferation and

migration. Following activation, a resolution phase occurs that involves basement membrane

reconstitution, inhibition of endothelial cell proliferation and migration, and stabilization of

newly formed vessels [33]. Migration and sprouting of endothelial cells is speculated to require

an endothelial-mesenchymal transition (EndMT) that must be reversed during the resolution

phase [34]. Impaired angiogenesis, resulting in reduced tumour MVD, can result in necrosis or

apoptosis of tumour cells as nutrient and oxygen delivery is limited [27].

Apoptosis is a regulated and ATP-dependent process of programmed cell death, whereas

necrosis is a pathological mechanism of cell death. Necrosis results from energy depletion and

elicits an immune response due to the rupture of necrotic cell membranes resulting in release of

cytoplasmic debris containing pro-inflammatory cytokines. Contrastingly, apoptosis is

infrequently accompanied by an inflammatory response as apoptotic bodies are phagocytosed

and cell membranes are not ruptured [35, 36]. Inflammation associated with necrosis can be both

tumour suppressing, by initiating an immune response that causes the destruction of tumour

cells, or tumour promoting, by increasing reactive oxygen species (ROS) and cytokine release

from infiltrating immune cells that induce DNA damage and replicative stress of tumour cells

[37]. Due to the multiple effects of necrosis-induced inflammation, the effect of necrotic tumour

area on prognosis is variable. For example, retrospective analysis of ovarian cancer patients that

received neo-adjuvant chemotherapy prior to debulking surgery revealed that absent or minimal

tumour necrosis was predictive of recurrent disease and prognosis [38]. Conversely, a correlation

between decreased necrotic tumour grading and improved prognosis was found in colorectal

patients [39]. Inflammatory signalling also has a role in promoting ascites formation that is

commonly identified in ovarian cancer patients.

Page 21: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

8

A majority of patients (80%) are diagnosed when metastases are present in the abdomen

or beyond the peritoneal cavity [2, 3]. Extensive seeding of the peritoneal cavity by tumour cells

in advanced stage disease is often accompanied by ascites, an accumulation of peritoneal fluid.

Ascites results from the excessive infiltration of fluid into the peritoneal cavity due to an

inflammation-dependent increase in vessel permeability. Furthermore, fluid accumulates due to

obstruction of lymphatic vessels by tumour cells causing an impairment of peritoneal fluid

drainage [40, 41]. Ascites is present in 77% of ovarian cancer patients at diagnosis and almost all

patients develop ascites during tumour progression [42]. Accumulation of ascites fluid induces

significant morbidity by increasing abdominal pressure that results in pain, early satiety,

respiratory compromise, nausea, and reduced mobility [40, 41]. Therapeutic options commonly

include paracentesis and diuretics followed by peritoneovenous shunts [40]. Interestingly,

peritoneovenous shunts used to drain peritoneal fluid from the peritoneum into the venous

circulation does not result in disseminated hematogenous metastases due to the unique

preference of ovarian cancer cells to the mesothelial layers of the peritoneum [28]. This form of

treatment is effective in relieving symptoms in approximately 70% of patients with high volume

ascites, although a subset of patients are resistant and ascites commonly recurs [41].

Furthermore, 6% of patients suffer major complications including pulmonary edema, pulmonary

embolus, and infection [41]. As there are limitations in clinical management of ascites, there are

further difficulties in treating ovarian cancer itself.

The current standard of care involves debulking surgery and intraperitoneal

administration of chemotherapeutics, allowing direct delivery to malignant cells [28]. Optimal

debulking results in minimal residual tumour lesions greater than 1 cm and is associated with

improved prognosis [28]. Due to the high incidence of acquired chemoresistance, 90% of

patients will succumb to this disease [2] which underscores the need to establish the initiating

events of each subtype in order to develop novel early stage screening approaches. The theory of

pathogenesis for HGSC, the most common and lethal ovarian cancer subtype, involves ovulation;

however, the actual cell affected is debated.

1.2.4 Theory of High-Grade Serous Carcinoma Pathogenesis

Ovulation is a physiological process initiated by the mid-cycle surge of luteinizing

hormone that induces rupture of the ovarian wall adjacent to a mature follicle and the ultimate

Page 22: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

9

release of an oocyte. This process involves necrosis, immune cell infiltration, vascular collapse,

and exposure of nearby cells to a transient pro-inflammatory environment [43]. In particular,

tissue repair following the liberation of the ovum results in chemokine release from cells of the

follicle that promotes immune cell recruitment [44]. Immune cells further maintain a pro-

inflammatory and oxidative environment by releasing cytokines and ROS. ROS production is

imperative for cumulus expansion and inducing cell signalling modifications that are crucial for

ovulation to occur; however, excessive ROS exposure can result in DNA damage [43, 45]. As

this is a physiological process that occurs repeatedly throughout a woman’s reproductive years, it

is vital that anti-inflammatory and tumour suppressor signalling mechanisms within exposed

cells are functional and inhibit malignant transformation. Inflammation and oxidative stress

associated with ovulation have a well-defined role in inducing DNA damage that could initiate

HGSC; however, exactly which cells are affected is debated.

There are two major proposed sites of origin of HGSC- ovarian surface epithelial (OSE)

cells and fallopian tube epithelial (FTE) cells [16]. The ovarian surface epithelium is a single

layer of uncommitted mesothelial cells that are protected from the intraovarian environment by

the tunica albuginea [16]. During the peri-ovulatory period, OSE cells are exposed to

inflammatory cytokines and ROS present in the follicular fluid which can cause DNA damage.

As a result of repair following ovulation, or as a result of aging, OSE cells can form inclusion

cysts in the hormone- and growth factor-rich ovarian stroma which induces replicative stress [16,

17]. These cells are predicted to be highly susceptible to malignant transformation as their cell

signalling systems are continuously disrupted due to mutations [17].

Transformed OSE cells have been shown to develop HGSC tumours. Mutation of the

retinoblastoma 1 (RB1) gene in normal OSE cells results in the development of ovarian tumours

that are characteristic of stage I HGSC tumours in mice [46]. Moreover, by additionally mutating

TP53, in the presence or absence of BRCA1 or BRCA2 mutations, tumours advance to stage IV

HGSC with associated peritoneal carcinomatosis. Gene expression analysis indicate that triple

mutant tumours have significant expression profile overlaps with human HGSC, supporting the

ability of transformed OSE cells to form HGSC tumours [46].

An additional theory supporting an OSE cell site of origin, independent of ovulation, is

that a population of OSE cells in the hilum region of the mouse ovary express stem cell markers

Page 23: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

10

and are susceptible to malignant transformation. Inactivation of transformation-related protein

53 (Trp53) and RB1 in these cells induces HGSC formation in vivo [47]. OSE stem cells are not

fully differentiated which may account for the Mullerian phenotype seen in cortical inclusion

cysts that is not present in fully differentiated OSE cells.

An important characteristic of cortical inclusion cysts present in ovaries is that they are

composed of Mullerian-like epithelium. OSE cells originate from a region of coelemic

epithelium overlying the gonadal ridge during development and are mesothelial. Studies propose

that when OSE cells are exposed to the new environment of the stroma during cyst development,

cells acquire a Mullerian epithelial phenotype as they undergo metaplasia. These changes include

alterations in cell shape, secreted products, and epithelial markers [28, 48]. Importantly, FTE

cells compose a Mullerian epithelium that originates from an invagination beside the gonadal

ridge, separate from the ovarian surface epithelium [28]. The hypothesis that Mullerian epithelial

cells comprising ovarian cortical inclusion cysts may actually originate from the fallopian tube

cannot be excluded at present.

More recently, the fallopian tube has emerged as the possible site of origin of HGSC. It

has been postulated that as fimbria are in close proximity to the ovary, distal FTE cells are

exposed to the same peri-ovulatory inflammatory and oxidative microenvironment. Excessive

oxidative stress associated with ovulation induces DNA damage that can result in an increased

expression of mutated p53 as evident by a “p53 signature” on immunostained tissue sections.

The combination of DNA damage and loss of cell cycle regulation by p53 promotes clonal

expansion of these cells and the development of tubal intraepithelial carcinomas [16]. These

lesions can then progress into invasive serous carcinoma. Invasive cells exfoliate from the

fimbria and implant on the ovarian surface and the peritoneum nearly simultaneously [21]. There

is compelling evidence supporting a fallopian tube site of origin, including the ability of FTE

cells to transform into HGSC.

Analysis of fallopian tube epithelium post-ovulation in vivo reveals that a pro-

inflammatory microenvironment, mediated by macrophage infiltration, is created. As a

repercussion of peri-ovulatory oxidative stress, FTE cells harbour DNA damage that may result

in signalling disruption and ultimate transformation of these cells [49]. Experimentally, normal

FTE cells have the capability to be transformed and ultimately create HGSC tumours in vivo.

Page 24: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

11

Karst et al. [50] transformed immortalized fallopian tube secretory epithelial cells by oncogenic

activation of myelocytomatosis oncogene (c-Myc), a locus commonly amplified in HGSC, and v-

Ha-ras Harvey rat sarcoma viral oncogene homolog (H-Ras), a constitutively active GTPase

that activates the mitogen-activated protein kinase (MAPK) proliferation pathway. Transformed

FTE cells were injected into the peritoneum of female Nude mice and gross examination

revealed that these cells had increased proliferative and metastatic capacities indicated by

widespread metastases throughout the abdomen. Importantly, tumour histology was consistent

with that of poorly differentiated HGSC [50]. In an additional study, Kim et al. [15]

demonstrated that tissue-specific double knockout of dicer 1 ribonuclease type III (DICER1) (an

essential gene in microRNA synthesis) and PTEN (a tumour suppressor that negatively regulates

the PI3K pathway) in the fallopian tube of mice resulted in HGSC tumour and ascites

development. Microarray analysis of primary tumours indicated a significant overlap in gene

expression with human HGSC [15]. An additional approach to support a fallopian tube site of

origin for ovarian cancer involves gene expression analysis of human tumours.

Gene profiling of RNA from human EOC tumours reveals that HGSC tumours

significantly correlate with fallopian tube epithelium brushings, rather than ovarian surface

epithelium. Interestingly, other common EOC tumour subtypes- clear cell, endometrioid, and

mucinous- did not have significant gene expression overlap with OSE brushings. This suggests

that only germ cell and sex cord-stromal tumours may actually originate from the ovary [51]. To

investigate this similarity of FTE cells with HGSC, Tone et al. [52] evaluated fallopian tube

epithelium of patients who were at increased risk of developing ovarian cancer due to inherited

BRCA1/2 mutations. Gene expression analysis revealed that FTE cells of the luteal phase in

BRCA mutation carriers had similar gene expression to HGSC [52]. In a follow-up study it was

identified that BRCA1-luteal phase FTE cells had elevated pro-inflammatory signalling and

altered response to ovulation-associated cytokines [53]. This gene expression profile would limit

inflammation mediation during the peri-ovulatory period in BRCA mutation carriers and may

contribute to the mechanism by which these patients are at increased risk of developing HGSC

[53].

Due to the increased risk for developing ovarian cancer, BRCA mutation carriers are

provided with the option to undergo prophylactic bilateral salpingo-oophorectomy (removal of

Page 25: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

12

both the ovaries and fallopian tubes). Extensive analysis of fallopian tubes collected as a

preventative measure reveals important information as to the site of origin of HGSC. Medeiros et

al. [54] identified that of 13 BRCA mutation carriers undergoing a prophylactic bilateral

salpingo-oophorectomy, 38% had serous intraepithelial carcinomas in the fallopian tube, but not

in the ovary. Similarly, Callahan et al. [55] identified that 7 of 122 BRCA-positive patients had

early malignancy that originated in the fimbrial or ampullary region of the fallopian tube. Only

one of these cases had an ovarian implant [55]. Independent of BRCA status, a study analyzing

42 serous “ovarian” carcinomas collected during surgical cytoreduction found that 71% involved

the fallopian tube [55]. These data support a fallopian tube site of origin. The possibility that

HGSC originates from the fallopian tube has valuable clinical implications in terms of risk

reducing surgery as BRCA patients may only require prophylactic fallopian tube removal.

1.2.5 Theory of Clear Cell Carcinoma Pathogenesis

Similar to HGSC, clear cell carcinoma is thought to arise from extraovarian tissue. Clear

cell carcinoma differs from other histotypes as it is often diagnosed at early stages. Due to the

common occurrence of chemoresistance, clear cell carcinoma has the poorest stage I prognosis of

all EOC subtypes. Information regarding the pathogenesis of this subtype is limited. Gene

profiling of RNA from human EOC tumours reveals that clear cell tumours significantly

correlate with normal endometrium brushings, rather than OSE [51]. The proposed theory of

pathogenesis of clear cell carcinoma involves neoplastic transformation of endometriotic lesions

that transplant onto the ovarian and periotoneal surface [21]. Activation of oncogenic pathways

of endometriotic cells allows endometrial tissue to implant, survive, and invade ovarian and

peritoneal tissue. Support of an endometrial site of origin for clear cell carcinomas is provided by

epidemiological evidence that suggests a protective effect for tubal ligation, which prevents

passage of endometrial tissue to the ovary and peritoneum. This protective effect is not evident in

the mucinous and serous histotypes that are not thought to arise from endometrium tissue [21].

Furthermore, epidemiological data indicates women who have endometriosis or ovarian

endometriosis (transplantation of non-malignant endometrium) are at greater risk of developing

clear cell carcinoma, as compared to controls [56].

Endometriosis is a gynaecological condition whereby endometriotic cells lining the

uterus displace and transplant to ectopic locations including the peritoneum, ovary, or

Page 26: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

13

rectovaginal septum. The most common explanation for transplanted endometrial tissue, which

does not apply to all cases of endometriosis, involves retrograde menstruation causing

endometrial debris to relocate and implant at ectopic sites [21]. Retrograde menstruation contains

pro-oxidant factors such as heme and iron that can promote ROS production and ultimate DNA

damage. Malignant transformation of endometriosis may be attributable to DNA mutations

induced by oxidative stress [21]. Greater than 90% of women experience retrograde

menstruation; however, the immune system clears displaced endometrium tissue in most women

[57]. Factors that allow endometriotic tissue to survive and form endometriosis are poorly

understood but are thought to include genetic predisposition, toxins, and immune status [57].

There are several mutations that are commonly acquired in the clear cell histotype.

Inactivation of the tumour suppressor gene ARID1A is seen in approximately 50% of cases.

Activating mutations of the oncogene (PIK3CA) that ultimately increases tumour promoting

PI3K/Akt signalling pathway, is also seen in 50% of tumours [21]. Activation of mammalian

target of rapamycin (mTOR), a downstream effector of PI3K/Akt (v-akt murine thymoma viral

oncogene homolog 1, Akt; also known as protein kinase B), has also been identified in clear cell

carcinomas and is involved in promoting cell cycle progression, proliferation, and angiogenesis

[56].

Due to the unique molecular signature of clear cell tumours, their chemoresponse rate to

drugs currently used to treat all ovarian histotypes differs greatly from other histotypes. Recent

genomic profiling of clear cell primary tumours revealed activation of pathways that are

involved in hypoxic cell growth, angiogenesis, and glucose metabolism that are not characteristic

of other histotypes [58]. Chemoresistance of clear cell tumours is likely attributed to the

increased capacity of tumour cells to survive in an environment with limited nutrients and

oxygen. Furthermore, in vivo studies indicate clear cell tumours are sensitive to anti-angiogenesis

chemotherapeutics, whereas serous tumours were not, suggesting an effective clear cell

histotype-specific treatment option [58]. It has also been identified that clear cell tumours often

have lower proliferative indexes, as compared to serous tumours. Low proliferative index of

clear cell tumours is correlated with reduced chemosensitivity that is likely attributable to

compromised intracellular drug accumulation in resting cells [59]. As clear cell tumours appear

to be the most resistant of all histotypes at early stages, it is imperative to identify gene mutations

Page 27: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

14

in clear cell tumours that may account for this; Veph1 is a novel gene that is upregulated in copy

number in a subset of clear cell carcinoma cell lines [5].

1.3 Veph1

1.3.1 Discovery, Structure, Function

Vertebral ventricular zone expressed PH domain protein (Veph1) mRNA was discovered

in the ventricular zone of the murine fetal brain and has a critical function in the development of

the zebrafish central nervous system [12]. Prior to this discovery, the Drosophila orthologue of

Veph1, Melted, was shown to have an imperative role in the physiological development of the

Drosophila peripheral nervous system [60, 61]. Tissue expression data are limited. Muto et al.

[12] identified murine Veph1 mRNA in the developing embryonic eye and brain whereas adult

murine Veph1 tissue expression was restricted to the eye and kidney. Similar expression data

were found in zebrafish tissues [12].

There are four reported human transcript variants of Veph1 that encode 833 (full-length

Veph1; isoform 1), 788 (isoform 2), 213 (isoform 3), and 175 (isoform 5) amino acid isoforms.

Isoform 2 comprises the entirety of the full-length isoform 1, with an omission of amino acids

626-670. Isoforms 3 and 5 are C-terminus truncations of isoform 1 (Figure 1A). These Veph1

isoforms appear to be formed by alternative RNA splicing. The full-length human Veph1 protein

has 85% sequence homology with murine Veph1A with greatest sequence identity is within the

N- and C-terminal regions. In both human Veph1 and murine Veph1A, a conserved pleckstrin

homology (PH) domain is located from amino acid 717 to 821. A shorter murine Veph1

transcript, Veph1B, has been identified that encodes a 253 amino acid isoform corresponding

with the Veph1A C-terminal amino acids 580-833 containing the PH domain (Figure 1B).

Importantly, primers used in murine Veph1 expression studies did not distinguish between the

murine Veph1A and Veph1B isoforms [12]; expression of these individual isoforms has thus not

been investigated.

The PH domain was initially discovered in 1993 as an approximately 100 amino acid

residue motif that was present in several eukaryotic signal-transducing proteins. Proteins that

contain PH domains include serine/threonine protein kinases, tyrosine kinases, small G-protein

regulators, and phosphoinositide metabolizing enzymes [62]. Initially, the PH domain was

Page 28: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

15

Figure 1: Schematic representation of human and murine Veph1 protein isoforms. A)

There are four transcript variants of human Veph1 identified that encode 833 (full-length Veph1;

isoform 1), 788 (isoform 2), 213 (isoform 3), and 175 (isoform 5) amino acid isoforms. Isoform

2 comprises the entirety of the full-length isoform 1, with an omission of amino acids 626-670.

Isoforms 3 and 5 are C-terminus truncations of isoform 1. B) There are two transcript variants of

murine Veph1 identified that encode 833 (Veph1A) and 253 (Veph1B) amino acid isoforms.

Veph1B is a 253 residue isoform corresponding with the Veph1A C-terminal amino acids 580-

833 that contain the PH domain. In both human and murine Veph1 proteins, a conserved PH

domain is located from amino acid 717 to 821. Amino acid sequence identity for the mid-region

and N- and C-termini between human (hVeph1) and murine (mVeph1) are indicated.

Page 29: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

16

thought to function primarily as an anchor for PH domain host proteins to the cell membrane by

binding phospholipids. It has most recently been identified that cell membrane anchoring may

also be due to PH domain interactions with membrane receptors and that PH domains most

frequently serve as a protein-protein interaction platform. There is no specific cellular

compartment or function of PH motifs [62, 63]. As Veph1 contains a PH domain, it is likely that

this protein interacts with cell signalling mediators.

Barrios-Rodiles et al. [64] preformed a high-throughput protein-protein interaction screen

and identified the PH domain of murine Veph1A to be a novel interactor of TGF-β/BMP (bone

morphogenetic protein; BMP) signalling mediators including ALK2, ALK5, ALK6 (activin

receptor-like kinase 2/5/6; ALK2/5/6), and SMAD1 [64]. Furthermore, Teleman et al. [8]

discovered that Melted, the Drosophila orthologue of Veph1, formed PH domain-dependent

interactions with membrane phosphoinositides that resulted in cell membrane localization.

Melted recruited phosphorylated FoxO and tuberous sclerosis 1 (Tsc1) to the cell membrane

which prevented FoxO signalling and inhibition of mTOR signalling; interactions with these

proteins was PH domain-independent. mTOR signalling was promoted as Melted impeded

inhibitory Tsc1/2 complex formation [8]. Overexpression of Melted was found to promote tissue

overgrowth and this effect was PH domain-dependent [8]. Importantly, Melted and human

Veph1 proteins are highly conserved with 87% PH domain and 76% N-terminal sequence

identity thus suggesting similar functions of both proteins. Analysis of human Veph1 via the

Eukaryotic Linear Motif server (http://elm.eu.org) predicts putative Forkhead-associated domain

ligand domains (LIG_FHA_1, LIG_FHA_2) and 14-3-3 proteins interacting motifs in the Veph1

protein sequence (LIG_14-3-3_2, LIG_14-3-3_3). 14-3-3 proteins mediate FoxO signalling and

14-3-3ε was shown to interact with Melted [8]. The identified interaction of Veph1 with

members of the TGF-β/BMP cascades, and potential interaction with FoxO signalling mediators,

suggests that Veph1 may have a role in modulating these signalling pathways that have predicted

roles in ovarian tumourigenesis.

Our lab has recently generated microarray data comparing SKOV3 human ovarian cancer

cells that express low endogenous levels of Veph1 (SKOV3-M) and SKOV3 cells stably

transfected with Flag-tagged Veph1 cDNA (SKOV3-Ve). Analyses of these data indicate that

Veph1 expression alters downstream gene targets of the FoxO, TGFβ, and Hippo signalling

Page 30: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

17

pathways (Shathasivam et al., in preparation). Each of these pathways has important

physiological and tumourigenic roles. Establishing tissue expression and the effect of Veph1 on

tumour properties warrants investigation.

1.3.2 Veph1 Involvement in FoxO Signalling

FoxO proteins compose the ‘O’ subgroup of the forkhead box family of transcription

factors. These proteins primarily act as tumour suppressors by targeting genes involved in cell

cycle regulation, apoptosis, and antioxidant defense [65]. In the presence of oxidative stress, as

occurs during ovulation, FoxO proteins act to protect cells from malignant transformation. This

is accomplished by inducing cell cycle arrest, upregulating DNA repair proteins [66], and

promoting ROS detoxification by increasing superoxide dismutase 2 (SOD2) and catalase

enzyme expression [65]. Reduced activity of these enzymes can result in accumulated DNA

damage and cell signalling disruptions. When cells are exposed to excessive oxidative stress,

FoxO transcription factors promote apoptosis by upregulating the expression of pro-apoptotic

proteins- such as Bcl2-like protein 11 (Bim), Bcl2 binding component 3 (Puma), and B-cell

lymphoma 6 (Bcl-6)- and proteins involved in cell cycle regulation- such as p27kip1, p21cip1,

p15INK4B [65]. Thus, in response to peri-ovulatory inflammation, FoxO transcription factors

are important in preventing excessive proliferation and survival of cells that are at high risk for

acquired mutations [16, 17]. Conversely, the FoxO signalling pathway may also be involved in

later events of tumour progression by promoting angiogenesis and invasion.

FoxO transcription factors upregulate SOD2 proteins in response to oxidative stress via c-

Jun N-terminal kinases (JNK)-mediated phosphorylation and activation of FoxO proteins [67].

Proteomic analysis of SKOV3 ovarian cancer cells, compared to a subseries SKOV3.ip1 cells

acquired from ascites with higher metastatic capability, indicates that SOD2 protein expression is

correlated with increased metastatic potential [68]. The effect of SOD2 on invasion is in part due

to indirect upregulation of matrix metalloproteinase (MMP) via hydrogen peroxide (H2O2)

production. Increased expression of SOD2 and decreased expression of catalase, resulting in

increased H2O2, exacerbates the effect of SOD2 on invasion [69]. Whether this SOD2-mediated

increase in MMP production is due to FoxO-dependent regulation of SOD2 has yet to be

established. FoxO-dependent MMP production has been identified in vitro and FoxO nuclear

retention due to serum starvation results in increased production of MMP-9 and MMP-13 [70].

Page 31: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

18

Determining whether the effect of FoxO on MMP expression is SOD2-dependent was not

examined. Studies evaluating the link between FoxO and SOD2 on promoting invasion warrant

investigation.

Furthermore, H2O2 (produced in part by the enzymatic conversion of superoxide by

SOD2 [65]) has been shown to increase angiogenesis by regulating several H2O2-sensitive

signalling pathways [71]. For example, H2O2 exposure causes inhibition of PTEN, resulting in

activation of the PI3K/Akt signalling pathway that activates hypoxia-inducible factor 1 (HIF1α).

HIF1α activation promotes the transcriptional upregulation of potent pro-angiogenic vascular

endothelial growth factor (VEGF) production [71]. Whether SOD2 and/or FoxO proteins had a

role in upregulating H2O2 production in this study was not investigated. VEGF indirectly

increases SOD2 production by promoting the activation of nicotinamide adenine dinucleotide

phosphate (NADPH) oxidase, resulting in ROS production [71]. These data indicate a FoxO-

independent mechanism by which SOD2 is upregulated during angiogenesis. FoxO1 and FoxO3a

signalling represses expression of potent vasodilator and pro-angiogenic factor nitric oxide

synthase 3 (NOS3; also known as eNOS) [72]. Altogether, these data suggest that the role of

FoxO in angiogenesis is complex and warrants further investigation. Disruptors of FoxO

signalling will be valuable to investigate as this pathway can be both tumour suppressive and

promoting. Veph1 appears to be a novel potential disruptor of this pathway.

Gene Set Enrichment Analysis of microarray data comparing control SKOV3-M and

Veph1-overexpressing SKOV3-Ve cells indicates that Veph1 alters the expression of 331

putative FoxO4-regulated genes. For example, genes that are partly regulated by FoxO that were

altered by Veph1 expression are involved in angiogenesis- such as NOS3 [72]- and apoptosis-

such as B-cell lymphoma 2 (Bcl2) [65]. Inhibition of FoxO signalling by Melted, a Veph1

Drosophila orthologue, has been described by Teleman et al. (2005) [8]. Melted contains two

forkhead-associated domain ligand domains in its N-terminal region that allow it to directly

interact with and localize FoxO at the cell membrane, preventing FoxO activity. This study also

identified an interaction between Melted and 14-3-3ε proteins, which are mediators of FoxO

signalling [8] (Figure 2A). Co-immunoprecipitation experiments conducted in our lab reveal that

human Veph1 interacts with 14-3-3ε and FoxO3a proteins (Shathasivam et al., in preparation).

Page 32: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

19

Figure 2: Schematic illustrating proposed mechanisms by which Veph1 modulates FoxO,

TGF-β, and Hippo signalling. A) Tyrosine kinase receptor binding of growth factors initiates a

phosphorylation cascade ultimately activating the kinase Akt. Akt translocates into the nucleus

and phosphorylates FoxO transcription factors at three distinct sites. This initiates a complex

process that results in the removal of FoxO proteins from the nucleus thereby diminishing the

transcription of their target genes. Inactive FoxO transcription factors interact with 14-3-3

proteins that facilitate the retention of FoxO in the cytoplasm. Studies conducted in Drosophila

indicate that Melted, a Veph1 homolog, interacts with both 14-3-3 and FoxO proteins and

inhibits FoxO activity by promoting its cell membrane localization [8]. Due to significant

sequence homology between Melted and Veph1, it is possible that Veph1 serves to inhibit FoxO

activity and expression of its target genes by interacting with 14-3-3 and FoxO proteins at the

cell membrane. Red question marks indicate interactions that have yet to be validated in human

cells. B) Binding of TGF-β dimers to their receptors results in a phosphorylation cascade that

activates SMAD2/3. These phosphorylated proteins then bind SMAD4 creating a complex that

translocates into the nucleus to regulate TGF-β target genes. Veph1 has been identified as an

inhibitor of canonical SMAD2/3-dependent signaling. Veph1 impedes the dissociation of

SMAD2 from TGF-β receptor 1 thereby reducing the nuclear localization of SMAD2/3/4

complexes and the expression of TGF-β target genes (Shathasivam et al., in preparation). C)

Several inputs (indicated by multiple arrows) regulate activation of MST1/2 kinases that

phosphorylate LATS1/2. Activated LATS1/2 kinases phosphorylate tumour promoting

YAP/TAZ transcription factors resulting in their interaction with 14-3-3 proteins and nuclear

exclusion, thereby decreasing the transcription of YAP/TAZ target genes. Studies investigating

Drosophila photoreceptor differentiation indicate that Melted inhibits the expression of wts

(LATS1/2 homolog), thus promoting YAP/TAZ nuclear activity. Wts also inhibits the expression

of melted [9, 10]. Due to significant sequence homology between Melted and Veph1, it is

possible that Veph1 also serves to inhibit LATS1/2 expression (indicated by dashed red line),

thereby promoting the expression of YAP/TAZ target genes. Orange-filled circles indicate

phosphorylation.

Page 33: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

20

Page 34: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

21

Alterations in phosphorylation status and localization of FoxO have been implicated in

endometrial, breast, and thyroid cancer [73]. FoxO signalling interruption can be attributed to

over-activation of the PI3K/Akt pathway, resulting in a loss of FoxO transcriptional activity [73,

74]. Lu et al. (2012) have identified an overexpression of threonine-32 phosphorylated FoxO3a

in ovarian tumours that is associated with poor survival. Threonine-32 is an Akt-regulated

residue that, once phosphorylated, prevents FoxO3a activity by promoting its nuclear

localization [75]. Furthermore, decreased ovarian tumour expression of FoxO3a is significantly

associated with decreased overall survival [76]. A correlation between decreased FoxO activity

and poor survival indicates a tumour suppressing role of this signalling pathway. Recent data

supporting a tumour promoting effect of FoxO signalling indicates that nuclear FoxO3a has a

significant positive correlation with poor survival in patients with invasive breast carcinoma

[77]. A possible mechanism by which nuclear FoxO promotes an aggressive phenotype is via

crosstalk with the wingless-type (Wnt) signalling pathway. Tenbaum et al. [78] identified that

increased co-expression of FoxO3a and β-catenin induced a metastatic phenotype in colon cancer

cells which was inversely related with patient survival time [78]. These data indicate that FoxO

transcription factors have a role in ovarian tumourigenesis and that Veph1 may modulate this

signalling pathway.

An important finding described by Teleman et al. (2005) includes the effect of Melted on

mTOR signalling. Melted localized the mTOR-inhibiting Tsc1/2 complex at the cell membrane

via interactions with Tsc1, resulting in activation of mTOR signalling [8]. The mTOR signalling

pathway causes transcriptional upregulation of tumour promoting genes involved in

angiogenesis, proliferation, survival, and motility [8]. The PI3K/Akt signalling pathway, that

activates mTOR signalling, was the most frequently altered cancer related pathway in 93 primary

epithelial ovarian tumours and mTOR activation has been described in clear cell carcinomas [56,

79]. These data suggest Veph1 may have a role in the pathogenesis of clear cell carcinomas, due

to modification of mTOR signaling. Considering the significant homology between Veph1 and

Melted, it is possible Veph1 may be a modulator of both FoxO and mTOR signalling in ovarian

tumourigenesis.

Page 35: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

22

1.3.3 Veph1 Involvement in TGF-β Signalling

TGF-β is a multi-functional cytokine that has both tumour suppressing and promoting

roles. Protective effects of TGF-β signalling include suppressing tumourigenic inflammation and

regulating cytostasis and apoptosis [80]. TGF-β signalling has a protective, anti-inflammatory

role by inhibiting the development, proliferation, and function of immune cells which prevents

the tumourigenic effects of inflammatory responses [80]. In addition to this function, TGF-β

signalling maintains cell homeostasis by modulating expression of genes involved in cell cycle

regulation and apoptosis. For example, activated SMAD transcription factors form a

transactivation complex with FoxO proteins to increase the transcription of cyclin-dependent

kinase inhibitor p21CIP1 [81]. Despite the protective effects of TGF-β signalling, this pathway is

also involved in tumourigenesis.

Physiologically, TGF-β signalling promotes growth inhibition; however, tumour cells can

evade these protective effects by inactivating core components of this pathway or specific

abolition of the tumour suppressive effects of this pathway while maintaining signalling

responses that provide a tumourigenic advantage [24]. Tumour promoting functions of this

signalling pathway include regulation of invasion, EMT, angiogenesis, and immune response

evasion. TGF-β signalling has an established role in promoting metastases formation by

supporting EMT and invasion. This pathway initiates transcriptional upregulation of several

EMT-promoting transcription factors [82] including Snail which is responsible for suppressing

epithelial-specific genes and promoting the expression of genes associated with the

mesenchymal phenotype [83]. Slug, an additional TGF-β regulated transcription factor, and Snail

promote the EMT phenotype of ovarian cancer cell lines by inducing expression of mesenchymal

marker vimentin and repressing epithelial marker E-cadherin [82]. Haslehurst et al. [84]

discovered that knockdown of Snail and Slug resensitized ovarian cancer cells to cisplatin

suggesting that TGF-β signaling may have additional roles in chemoresponse [84]. TGF-β

treatment also promotes MMP-dependent invasion of epithelial ovarian cancer cell lines which is

required following EMT of metastatic cells [85].

In addition to promoting metastases, TGF-β signalling is implicated in ascites formation

and angiogenesis. Orthotopic injection of human ovarian cancer cells with and without soluble

TGF-β receptor II treatment to inhibit TGF-β signalling revealed that TGF-β blockade inhibited

Page 36: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

23

ascites fluid accumulation by preventing abnormalization of the diaphragm lymphatic vessel

network allowing for improved ascites drainage. Furthermore, TGF-β signalling inhibition

resulted in a decrease of VEGF expression that ultimately reduced the prevalence of cluster of

differentiation 31 (CD31; also known as platelet endothelial cell adhesion molecule 1, PECAM-

1) positive endothelial cells [40]. In addition, TGF-β also promotes EndMT which may be an

important process in endothelial cell sprouting. Treatment of murine microvascular endothelial

cells with TGF-β increases the expression of the EMT-mediator Snail, resulting in EndMT

evident by morphology analysis and expression of epithelial and mesenchymal markers [86]. The

process of EndMT has also been implicated in the generation of cancer-associated fibroblasts

(CAFs) that contribute to tumourigenesis by releasing tumour promoting factors such as TGF-β

and VEGF, thereby further promoting angiogenesis [87]. As previously mentioned, TGF-β

signalling is immunosuppressive; however, when the immunosuppressive effects of TGF-β

outweigh the tumour-suppressive benefits of its anti-inflammatory action, a consequential net

protumourigenic advantage may result due to immune surveillance evasion by tumour cells [80].

Due to the pleiotropic effects of TGF-β signalling, up- and down-regulation of TGF-β signalling

has been identified in ovarian cancer samples.

Reports of contrasting effects on TGF-β signalling are evident in ovarian cancer samples.

For example, examination of ovarian tumours indicates that TGF-β isoforms are increased as

compared to benign tumours. Increased expression of the TGF-β3 isoform was significantly

associated with advanced stage disease [25]. Ovarian cancer studies have identified decreased

expression and disrupting mutations of TGF-β ligand binding receptors and downstream

signalling mediators [26]. Conversely, other studies have not identified inactivating mutations in

core mediators of this pathway in ovarian tumours [27]. Aberrant expression of positive and

negative modulators of TGF-β signalling has also been reported in ovarian tumours. For

example, microarray expression profiling reveals that TGF-β signalling inhibitors- such as BMP7

and dachshund homolog 1 (DACH1)- are up-regulated in advanced-stage ovarian serous

carcinomas [26]. In contrast, TGF-β signalling enhancers- such as enhancer of zeste homolog 2

(EZH2)- are also up-regulated [28]. These data suggest that TGF-β signalling is disrupted in

ovarian cancer. Due to the pleiotropic effects of TGF-β signalling in tumour progression, it is

difficult to predict whether loss or gain of TGF-β signalling will potentiate disease progression.

Page 37: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

24

Identifying novel disruptors of this pathway and establishing the net effects of each modulator on

tumourigenesis is imperative.

Veph1 has recently been identified as a negative signalling modulator of canonical TGF-

β signalling. Veph1 limits nuclear accumulation of SMAD2 by impeding the dissociation of

SMAD2 from TGF-β receptor 1 (TβRI) thus reducing the transcription of TGF-β regulated genes

(Shathasivam et al., in preparation; Figure 2B). Genes affected by Veph1 expression were

identified in a microarray analysis of SKOV3 human ovarian cancer cells (Shathasivam et al., in

preparation). Veph1 increased expression of negative modulators of TGF-β signalling- such as

transmembrane prostate androgen-induced protein (TMEPAI) and BMP and activin membrane-

bound inhibitor homolog (BAMBI)- suggesting a possible SMAD-independent mechanism of

Veph1-mediated inhibition of TGF-β signalling. Veph1 caused decreased repression of the

epithelial marker CDH1 (gene encoding the epithelial marker E-cadherin) by TGF-β. Increased

E-cadherin would impair the EMT-promoting activity of TGF-β signalling and suggest Veph1

may be present in epithelial cells. Veph1 reduced the expression of potent pro-angiogenic factor

VEGFA mRNA which suggests it may inhibit angiogenesis as well. Interestingly, microarray

data indicated that positive modulators of TGF-β signalling- such as betaglycan (TGFBR3) and

disabled homolog 2 (DAB2)- were also up-regulated in the presence of elevated Veph1 levels

indicating that Veph1 may promote non-canonical TGF-β signalling. As TGF-β signalling has

pleiotropic effects, it is imperative to determine the ultimate effect of Veph1 on TGF-β-regulated

disease progression.

It is important to note that Veph1 has been implicated as possibly promoting EMT of

breast cancer cells [88]. Wnt signalling enhances the metastatic capability of breast cancer cells

and this may be mediated in part by Veph1. Breast tumour-associated fibroblasts secrete

exosomes that are internalized by breast cancer cells and induce Wnt-planar cell polarity

signalling that is necessary for protrusive activity, motility, and metastases. Proteomic analysis

reveals that Veph1 is present in these secreted exosomes, suggesting a possible role of Veph1 in

Wnt-induced metastasis [88]. The specific role of Veph1 in this study was not examined [88].

Microarray data from our lab indicates Veph1 induces up-regulation of Wnt ligands and nuclear

mediators of canonical Wnt/β-catenin signalling- such as lymphoid enhancer-binding factor 1

(LEF1) and transcription factor 4 (TCF4). This increase in Wnt signalling mediators translated

Page 38: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

25

to increased Wnt signalling, as determined by a reporter assay measuring LEF/β-catenin

transcriptional activation in vitro (Shathasivam et al., in preparation). By promoting Wnt

signalling, Veph1 may support tumourigenesis.

1.3.4 Veph1 Involvement in Hippo Signalling

The Hippo tumour suppressor signalling pathway has a key role in regulating target genes

involved in cell proliferation and apoptosis to control tissue growth and organ size. Hippo

signalling involves activation of mammalian STE-20-like kinase 1/2 (MST1/2) kinases that act to

phosphorylate and activate LATS1/2 kinases which phosphorylate transcription factors yes-

associated protein (YAP) and transcriptional coactivator with PDZ-binding motif (TAZ),

resulting in their inhibition via nuclear exclusion [89]. Increased YAP expression induces

tumourigenic activity of human and Drosophila OSE cells by regulating gene expression that

promotes proliferation, reduces apoptosis, and increases cell migration and anchorage-

independent growth [90]; similar findings have been identified in clear cell carcinoma cells with

increased expression of YAP [91]. YAP signalling is also proposed to promote metastasis and

angiogenesis. Huntoon et al. [92] discovered that in vitro and in vivo inhibition of LATS1 and

LATS2 in ovarian cancer cells and primary tumours resulted in increased YAP activity evident

by accumulation of the target protein connective tissue growth factor (CTGF) that promotes cell

migration, proliferation, and angiogenesis [92]. Furthermore, YAP expression has been identified

as critical for the establishment and maintenance of CAFs that induce matrix remodeling

required for angiogenesis [93]. The Hippo signalling pathway crosstalks with additional

pathways by activating FoxO transcriptional activity during oxidative stress, and inhibiting TGF-

β and Wnt signalling [89]. As the Hippo pathway interacts with several signalling cascades,

modulators of this pathway have widespread implications. Dysregulation of the Hippo signalling

cascade has been identified in ovarian cancer.

Analysis of 284 human EOC samples reveals that approximately 15% of tumours express

elevated levels of activated nuclear YAP expression. High levels of nuclear YAP correlated with

a significant reduction in progression free survival in ovarian cancer patients [91]. These data are

further supported by an additional independent analysis of 70 human primary tumours that

revealed nuclear accumulation of YAP and reduced cytoplasmic retention of phosphorylated

YAP is correlated with approximately 50% lower 5-year survival and promotes resistance to

Page 39: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

26

cisplatin-induced apoptosis in OSE cells [90]. Increased YAP expression has also been shown to

promote EMT, migration, anchorage-independent growth, and chemoresistance to cisplatin and

taxol in clear cell carcinoma cell lines [91]. Identifying novel Hippo signalling modulators that

promote YAP activity will be valuable to develop potential novel therapeutic targets.

Microarray data analyzed by our lab reveals that Veph1 expression has an impact on 274

YAP-regulated genes, suggesting that Veph1 expression impacts Hippo signalling. Whether

these gene expression alterations cause a net pro-tumourigenic or protective effect is yet to be

examined. Veph1 involvement in Hippo signalling is supported by studies conducted in

Drosophila that suggest Melted (Veph1 homolog) negatively regulates the transcription of the

wts (LATS1/2 homolog) gene, resulting in the differentiation of R8 photoreceptors to express

blue-sensitive rhodopsins (Figure 2C). Furthermore, Wts inhibits melted expression resulting in

the differentiation of R8 photoreceptors to express green-sensitive rhodopsins [9-11]. Due to the

significant homology between Veph1 and Melted, these data suggest Veph1 may be involved in

preventing tumour suppression by Hippo signalling within human cells.

1.3.5 Veph1 Expression in Epithelial Ovarian Cancer

The gene encoding human Veph1 is located on chromosome 3 at locus q24-q25 which

has been implicated as a site of mutation in tumour samples. A mutation discovery screen of

breast cancer samples indicates there was a significant frequency of VEPH1 mutations [94];

furthermore, an additional breast cancer genomic data analysis discovered an increase in VEPH1

copy number in breast tumour tissue as compared to normal tissue [95]. Copy number gain of a

region encompassing the VEPH1 locus has also been identified in patients with gastrointestinal

stromal cancer and may be correlated with poor survival [96]. In addition to these cancers,

mutations in VEPH1 and its locus region have been reported in ovarian cancer.

Array-based comparative genomic hybridization analysis preformed on 12 human

ovarian clear cell carcinoma cell lines found there was an amplification of the 3q24-q26 region,

encompassing the VEPH1 locus, in 7 of these cell lines [5]. Recent genome-wide association

studies of ovarian cancer patients and controls identified the 3q25 region as a susceptibility locus

that increases the risk of developing EOC [97]. Furthermore, an independent genome wide

analysis study of copy number in 68 primary human epithelial ovarian carcinomas indicated that

Page 40: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

27

VEPH1 copy number and mRNA expression was significantly increased in >40% of these

samples [4]. These data raise the possibility that VEPH1 could be a novel ovarian cancer driver

gene, and may specifically be involved in the clear cell histotype.

Further support of a possible tumour promoting function of Veph1 involves

characterization of human ovarian cancer cell lines. The ES2, HEY, and OVCA429 ovarian

cancer cell lines are highly motile and exhibit invasive properties due to their ability to degrade

cross-linked collagen type I via membrane-type I matrix metalloproteinase (MT1-MMP)

upregulation [6, 7]. These aggressive cell lines have high levels of Veph1 protein as compared to

SKOV3 and OVCAR3 ovarian cancer cell lines that have non-detectable Veph1 levels

(Shathasivam et al., in preparation) and have a less invasive phenotype [6, 7]. HEY, ES2, and

OVCA429 ovarian cancer cell lines also form compact multicellular spheroids aggregates that

invade 3-D matrix culture systems while SKOV3 and OVCAR3 cell lines form loose aggregates

[7]. As compact spheroids exhibit enhanced resistance to anoikis following treatment with

chemotherapeutic agents, it is possible Veph1 expression status may modulate chemoresponse

[30, 31]. These data suggest that Veph1 expression may have a role in supporting aggressive

behaviour and chemoresistance of ovarian cancer cells by promoting invasion and compact

spheroid formation.

1.4 Thesis Hypothesis and Objectives

Veph1 may be a novel modulator of pathways involved in ovarian tumourigenesis. Our

lab has recently generated microarray data indicating Veph1 expression alters genes that are

partially regulated by the FoxO, TGF-β, and Hippo signalling pathways, which have both tumour

suppressing and promoting functions (Shathasivam et al., in preparation). These data are

supported by studies conducted in human and Drosophila cells [8-10] . A role for Veph1 in EOC

is strongly implicated by genomic analyses indicating an increase in VEPH1 gene copy number

in a subset of ovarian primary tumours and cell lines [4, 5]. Furthermore, Veph1 expression

correlates with an invasive and motile phenotype in ovarian cancer cell lines [6, 7].

Determination of the impact of Veph1 on ovarian tumourigenesis is unexplored but certainly

warranted. Given that only a single publication has addressed the expression of Veph1 in a

mammalian species, it is also important to confirm the reported expression of Veph1.

Page 41: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

28

Based upon papers indicating a role of Veph1 in modulating multiple signaling pathways,

I hypothesized that Muto et al. [12] may have underestimated the range of tissues expressing

Veph1. Expression data are relevant to studies in cancer as it is important to determine if levels

of Veph1 in malignant cells deviates from that of normal tissues or cells of origin. Expression

information may also suggest involvement in particular physiological processes. I further

hypothesized that Veph1 expression in ovarian cancer cells promotes tumour progression. Thus,

there were two specific objectives of this thesis. The first was to identify expression of Veph1A

and Veph1B mRNA and protein in murine embryos and adult tissues. As Veph1 may be

implicated in modulating several physiological signalling pathways, I expected Veph1

expression to be widespread. As microarray studies in our lab have indicated that Veph1

increases the expression of the epithelial marker E-cadherin, at both the RNA and protein level, I

anticipated that Veph1 protein expression would be predominantly expressed in epithelial cells

of tissues examined.

The second objective was to determine the effect of Veph1 expression on ovarian tumour

progression, using a xenograft mouse model. As VEPH1 gene copy number is increased in select

ovarian samples [4, 5], suggesting a possible driver function, and Veph1 protein expression is

correlated with an aggressive phenotype of ovarian cancer cell lines [6, 7], I predicted Veph1

would promote tumour progression. I examined tumour growth, proliferative index, necrotic

tumour area, and microvessel density to investigate the effect of Veph1 expression on primary

tumours.

Page 42: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

29

2.2. Chapte

MATERIALS AND METHODS

2.1 Tissue Culture

SKOV3 human epithelial ovarian cancer cells that express low endogenous Veph1 levels

were previously stably transfected with either a metallothionein-inducible vector containing

Flag-tagged Veph1 (SKOV3-Ve) or an empty vector (SKOV3-M) and were clonally selected

(Shathasivam et al., in preparation; Figure 3). The SKOV3-Ve3 clone was used for in vitro

experiments while the SKOV3-Ve4 clone was used for in vitro and in vivo experiments. Cells

were maintained as monolayers in Dulbecco’s Modified Eagle Medium: Nutrient Mixture F-12

(Invitrogen, Burlington, ON, Canada) (without phenol red) pH 7.4, supplemented with 0.25%

(v/v) fungizone (amphotericin B) (Invitrogen, Burlington, ON, Canada), 1% (v/v) penicillin

(5000 U/ml) (Invitrogen), 5% (v/v) fetal bovine serum (FBS) (HyClone Laboratories Inc.,

Logan, UT, USA), and geneticin (G418; 100 µg/ml) (Invitrogen). Culture medium was changed

every three days and cells were subcultured at 80% confluency. The cells were grown in a

humidified 37ºC incubator, in the presence of 5% CO2.

2.2 Tissue Specimen Collection

Murine adult tissue specimens were collected from 6 to 8 week old male and female ICR

mice, after euthanization by CO2 inhalation and cervical dislocation. Mice were provided by the

Centre for Cellular and Biomolecular Research Animal Facility (Toronto, ON, Canada). Whole

ICR murine embryos were obtained as previously described [98]. Specimens were snap-frozen in

liquid nitrogen within 2 minutes of excision and stored in -80ºC.

2.3 Animals

Twenty-four female (12 mice per treatment group), 5 to 8 week old, BALB/c nude mice

were purchased from Charles River Laboratories (code 194; Wilmington, MA, USA). Mice were

housed at the Centre for Cellular and Biomolecular Research Animal Facility (Toronto, ON,

Canada). All mice were maintained in a controlled and sterile environment with a 12-hour light-

12-hour dark cycle and had access to food and water. All mice received drinking water

Chapter 2

Page 43: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

30

Figure 3: Illustration of the pMEP5-3xFlag-Veph1 expression vector. Illustration of the

vector stably transfected into SKOV3-Ve cells. The pMEP5 vector was obtained from Dr. Jeff

Wrana, Mount Sinai Hospital. Veph1 cDNA was obtained from Open Biosystems (Thermo

Fisher Scientific; accession number BC101660). The 3xFlag-Veph1 cDNA was subcloned into

the pMEP5 vector at NheI and BamHI cloning sites. SKOV3-M cells were stably transfected

with the empty (omission of 3xFlag-Veph1 cDNA) pMEP5 vector.

Page 44: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

31

containing 25 mM zinc sulfate (ZnSO4) in order to induce Veph1 expression in SKOV3-Ve cells.

All procedures were performed in accordance with a protocol approved by the University of

Toronto Faculty of Medicine and Pharmacy Animal Care Committee.

2.4 Tumour Model

SKOV3-M and SKOV3-Ve4 cells were harvested by trypsinization at 80% to 85%

confluency and centrifuged at 250 g for 5 minutes at room temperature and washed twice with

1X-phosphate buffered saline (PBS). Cells were resuspended in 1X-PBS and counted three times

with a TC10 Automated Cell Counter (Bio-Rad Laboratories Inc., Hercules, CA, USA). Counts

were averaged and a total of 3 x 106 cells were suspended in 300 µl of 1X-PBS for tumour

inoculation of each mouse and placed on ice until injection. Following a 7-day acclimation

period, cells were injected subcutaneously into the flank of each mouse using a 26.5 gauge

PrecisionGlide® needle (Becton Dickinson & Co., Franklin Lakes, NJ, USA), approximately 30

minutes after harvesting. Twelve mice received SKOV3-M cells and the remaining 12 mice

received SKOV3-Ve4 cells. To ensure an accurate transfer of cells, mice were anaesthetized

using 2% to 5% isoflurane at 1 L/min O2 flow for 2 minutes. The volume of the resulting

xenograft tumours were recorded using a Traceable® digital caliper (Thermo Fisher Scientific,

Rockford, IL, USA) every 24 to 48 hours. Tumour volumes were calculated using the following

formula: π/6(length x width x depth) [99]. Animals were weighed once a week and general

health was monitored every 24 to 48 hours.

Animals were euthanized by CO2 inhalation and cervical dislocation once tumour

volumes reached 520 mm3 (as instructed by the facility veterinarian) or 63 days post-injection if

tumours did not form. The tumours were excised, weighed, and cut in half; one half was snap-

frozen and the other half was placed in 10% formalin fixative (Sigma, Oakville, ON, Canada) for

paraffin embedding.

Xenograft tumours were fixed overnight at 4°C in 10% formalin and washed for 1 hour

with 1X-PBS (pH 7.4), overnight with 70% ethanol at 4°C, 1 hour with 80% and 90% ethanol

each, and two 1 hour washes with both 100% ethanol (ethanol from J.T. Baker, Mississauga,

ON, Canada) and xylene (Thermo Fisher Scientific). All washes were performed at room

temperature. Specimens were incubated in heated paraffin overnight and then embedded.

Page 45: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

32

Embedded tissues were then sectioned (5 µm thick), using a Leica RM2255 microtome (Leica

Biosystems, Buffalo Grove, IL, USA).

2.5 RNA Extraction and Reverse Transcriptase-Polymerase Chain

Reaction (RT-PCR)

Mouse tissue specimens were mechanically homogenized, using a Caframo homogenizer

(Wiarton, ON, Canada), in Trizol (Ambion, Carlsbad, CA, USA) at 4°C. Total RNA was

extracted as specified by the manufacturer. RNA concentration was determined by measuring

absorbance at 260 nm with a spectrophotometer.

Three micrograms of total RNA were reverse transcribed using Oligo d(T)16 (Roche

Applied Science, Laval, QC, Canada) and SuperScript™ II reverse transcriptase (Invitrogen) as

per manufacturer’s protocol. The resulting cDNA was quantitated spectrophotometrically to

ensure equivalent amounts of cDNA were used per sample during amplification (1.8 µg for

whole embryos and non-reproductive tissue Veph1 isoforms; 2.5 µg for reproductive tissue

Veph1 isoforms; 300 ng for whole embryos and non-reproductive tissue actin; 400 ng for

reproductive tissue actin). cDNA was amplified using specific primers that cross intron-exon

boundaries (summarized in Table 1; illustrated in Figure 4). PCR was performed using Taq

polymerase (Fermentas, Burlington, ON, Canada) and involved: pre-incubation for 5 minutes at

95°C, denaturation at 94°C for 90 seconds, primer annealing at temperatures (annealing

temperature; Ta) indicated in Table 1 for 75 seconds, and extension at 72°C for 2 minutes. This

was repeated for 32 cycles and the resulting PCR products were resolved by 2% agarose gel

electrophoresis and subsequently verified by automated sequencing. Each panel of tissues was

repeated 3 times with specimens from different mice.

2.6 Protein Extraction and Western Blot Analysis

To confirm Veph1 induction via cadmium chloride (CdCl2; Sigma) treatment, 40,000

SKOV3-Ve3 cells were seeded per well in a 12-well plate with medium and were maintained in

the incubator for 24 hours to allow attachment. Cells were then treated with 1 µM CdCl2 for 24

or 48 hours. To examine SOD2 protein expression, 50,000 SKOV3-M, SKOV3-Ve3, and

SKOV3-Ve4 cells were seeded per well in a 24-well plate with medium and then treated with 1

µM CdCl2 for 24 hours (a 48 hour time point for SKOV3-Ve4 cells was included) to induce

Page 46: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

33

Table 1: Primers used for RT-PCR analyses of Veph1A, Veph1B, and ACTB.

Name Accession

number

Sequence (5'-3') Ta

(°C)

Amplicon

Size (bp) Veph1AFwd NM_145820 GAATACCAAGATAAGCTCTACTT 52 871

Veph1ARev ATGTGGTTACTTCTCGGCTTT

Veph1BFwd NM_028357 TTGCAGCCTTCTGGTGATCA 54 871

Veph1BRev ATGTGGTTACTTCTCGGCTTT

ACTBFwd NM_007393 AGTGTGACGTTGACATCCGTAAAGA 56 296

ACTBRev AGCTCAGTAACAGTCCGCCTAGAA

Page 47: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

34

Figure 4: Schematic illustrating RT-PCR primer location within Veph1A and Veph1B

transcripts. Illustration of the coding (blue) and non-coding (grey) regions of murine Veph1A

and Veph1B (region of sequence non-identity indicated by blue hatch lines). Primers used by

Muto et al., [12] are indicated with black arrows. Primers used in this study are indicated with

red arrows; the forward primers distinguish between Veph1A and Veph1B.

Page 48: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

35

Veph1 expression. Cells were harvested using Radioimmunoprecipitation Assay (RIPA) lysis

buffer (150 mM sodium chloride, 50 mM 4-2-hydroxyethyl-1-piperazineethanesulfonic acid

(HEPES; pH 7.4; Sigma), 1% (v/v) Triton X-100 (ICN Biochemicals Inc., Aurora, OH, USA)),

0.1% (v/v) sodium dodecyl sulfate (SDS; BioShop Canada Inc., Burlington, ON, Canada), 1%

(w/v) sodium deoxycholate (EMD Chemicals, Philadelphia, PA, USA), Complete Protease

Inhibitor Cocktail and 1X PhosSTOP Phosphatase Inhibitor tablets (both from Roche Applied

Science)) for 30 minutes at 4ºC. Each well was scraped and total cell lysates were collected and

centrifuged at 24,100 g for 15 minutes, at 4ºC. The supernatants, containing total protein extract,

were collected and stored at -80ºC.

For protein extraction from tissue specimens, tissues were mechanically homogenized in

RIPA buffer, containing protease and phosphatase inhibitor tablets (Roche), and centrifuged

twice at 24,100 g for 15 minutes, at 4ºC. The supernatants, containing total protein extract, were

collected and stored at -80ºC. Protein content of all the lysates was quantified with a

Bicinchoninic Acid (BCA) Protein Assay Kit (Pierce, Rockford, IL, USA) following the

manufacturer’s recommended protocol.

Aliquots of total protein were diluted in RIPA lysis buffer to obtain 8 µg to 40 µg per

well (as specified in figure legends). SDS loading buffer (1M Tris-HCl pH 6.8, 10% SDS, 5%

glycerol (BDH Chemicals, Mississauga, ON, Canada), 5% β-mercaptoethanol (Sigma) and

0.05% bromophenol blue dye) was added to the samples. Each sample was boiled for 5 minutes

and then loaded on to 8% to 12% SDS-polyacrylamide gels (as specified in figure legends) and

electrophoresis was performed. The proteins were transferred in transfer buffer (48 mM Tris, 39

mM glycine, 20% methanol; all from Thermo Fisher Scientific) to an activated Immun-blot

Polyvinylidene Fluoride (PVDF) membrane (Bio-Rad Laboratories Inc.) by electrophoresis at

125 mA, at 4ºC, for 1.5 hours. The membranes were then blocked for 1 hour, at room

temperature, in blocking solution comprised of 5% skim milk powder and 1% normal goat serum

(Vector Laboratories Inc., Burlingame, CA, USA) in 1XPBS-T (PBS containing 1% tween-20;

Sigma) and incubated overnight at 4ºC with primary antibodies diluted with blocking solution.

Primary antibody information is listed in Table 2. Membranes were then washed with 1XPBS-T

and incubated for 1 hour at room temperature with horseradish peroxidase (HRP)-tagged

Page 49: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

36

Table 2: Antibodies used for Western blot analyses.

Target Source Dilution Catalog Number Manufacturer

Veph1 Rabbit Polyclonal 1:100 HPA026645 Sigma

Flag Mouse Monoclonal 1:500 200472 Stratagene

SOD2 Rabbit Polyclonal 1:5000 ab13533 Abcam

GAPDH Rabbit Polyclonal 1:2000 sc-25778 Santa Cruz

Biotechnology Inc.

Page 50: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

37

secondary antibody targeting the host species of the primary antibody (1:1000; Santa Cruz

Biotechnology, Santa Cruz, CA, USA) and then washed twice with 1XPBS-T. Immmunoreactive

proteins were visualized by enhanced chemiluminescence (ECL; Santa Cruz Biotechnology)

following the manufacturer’s recommended protocol after exposure to HyBlot CL

Autoradiography Film (Denville Scientific, Inc., Woodbridge, ON, Canada). Immunoreactive

band intensities were quantified using Image Quant 5.0 software (Molecular Dynamics,

Pittsburgh, PA, USA). Each experiment was independently repeated three times, where

applicable. Data were normalized to glyceraldehyde 3-phosphate dehydrogenase (GAPDH)

levels.

2.7 Immunohistochemistry

Paraffin-embedded tissue sections were deparaffinised in xylene for 2X5 minutes and

rehydrated by washing with decreasing percentages of ethanol (100%, 95%, 70%) for 2X5

minutes each. Slides were then rinsed with double distilled water (DDW). Antigen retrieval was

performed using 10 mM citrate buffer, pH 6.0 (0.1 M citric acid and 0.1 M sodium citrate).

Sections were heated in a 750 W microwave oven at low setting for 4X5 minutes with gentle

agitation between heating intervals, followed by a 20 minute room temperature cooling period.

The slides were heated again for 5 minutes and cooled for 20 minutes [100].

Sections were then rinsed once with DDW and then 2X5 minutes with 1X-PBS. A 10

minute incubation, at room temperature, in 1X-PBS containing 0.3% (v/v) TritonX-100 was

performed to allow permeabilization. This was followed by 3X5 minute 1X-PBS washes at room

temperature. The slides were then incubated for 20 minutes, at room temperature, in 0.3% (v/v)

H2O2 (Sigma), dissolved in methanol (Thermo Fisher Scientific), and washed 3X5 minutes in

1X-PBS which was followed by a one hour incubation in 10% (v/v) normal goat serum in 1X-

PBS. Sections were incubated overnight at 4°C in a humidified chamber with either primary

antibody, or matching primary antibody host-species non-specific IgG (Santa Cruz

Biotechnology Inc.), diluted in 1X-PBS containing 1% normal goat serum. Primary antibody

information is listed in Table 3. Sections were then washed at room temperature for 3X5 minutes

with 1X-PBS prior to one hour incubation with biotinylated secondary antibody (Vector

Laboratories Inc.) targeting primary antibody IgG (diluted at 1:200 in 1X-PBS containing 1%

(v/v) normal goat serum). The slides were washed 3X5 minutes with 1X-PBS and then incubated

Page 51: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

38

Table 3: Antibodies used for immunohistochemical staining.

Target Source Dilution Catalog

Number

Manufacturer

Veph1 Rabbit Polyclonal 1:30 HPA026645 Sigma

PCNA Mouse Monoclonal 1:2000 M879 Dako

Ki-67 Mouse Monoclonal 1:200 0505 Immunotech

Cleaved Caspase-3 Rabbit Polyclonal 1:300 9661 Cell Signalling

CD34 Rabbit Monoclonal 1:250 ab81289 Abcam

CD31 Rabbit Polyclonal 1:50 *Outsourced ab28364 Abcam

SOD2 Rabbit Polyclonal 1:250 sc-25778 Abcam

Page 52: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

39

at room temperature with a 1:400 dilution of a streptavidin-peroxidase complex (Vector

Laboratories Inc.) in 1X-PBS, followed by 3X5 minute 1X-PBS washes.

To visualize stained protein, sections were incubated for 1-10 minutes with 3,3-

diaminobenzidine (DAB; 0.7 mg/ml in 1X-PBS; Sigma) in the presence of 0.03% (v/v) H2O2, at

room temperature. The sections were rinsed with DDW and counterstained, where applicable,

with Harris Hematoxylin solution (Sigma) for 40 seconds, followed by a 2 minute rinse with

running tap water. Sections were then dehydrated with increasing ethanol concentrations (70%

for 5 minutes, 95% and 100% for 10 minutes each) and then a 10 minute incubation in xylene.

Coverslips (VWR Scientific, Mississauga, ON, Canada) were then mounted with DPX permount

(Sigma). CD31 immunohistochemical (IHC) staining was performed by Pathology Services at

The Toronto Centre for Phenogenomics (Toronto, ON, Canada).

For hematoxylin and eosin staining, tumour sections were placed in hematoxylin for 40

seconds following rehydration of sections, and then rinsed for 2 minutes with running tap water.

Sections were then placed in Eosin Y (ICN BioChemicals Inc.) for 2 minutes and then rinsed for

2 minutes with running tap water. Slides were then dipped in 100% ethanol twice and incubated

in xylene for 10 minutes prior to mounting coverslips with DPX permount (Sigma).

2.8 Double Terminal Deoxynucleotidyl Transferase dUTP Nick

End Labeling (TUNEL) and Cleaved Caspase-3 Staining

Paraffin-embedded tumour sections were double stained for TUNEL, to detect dying

cells, and cleaved caspase-3, to detect apoptotic cells specifically. Each section was initially

stained for TUNEL using TACS2 TdT-DAB In Situ Apoptosis Detection Kit (Trevigen,

Gaithersburg, MD, USA) following the manufacturer’s protocol. Following completion of DAB

staining, sections were washed with 1XPBS for 20 minutes and then the Cell and Tissue HRP-

AEC Staining Kit was used to detect cleaved-caspase 3 using a 3-amino-9-ethylcarbazole (AEC)

red chromogen stain by following the manufacturer’s protocol. Sections were incubated with

cleaved caspase-3 primary antibody (Cell Signaling Technology, Danvers, MA, USA) at a 1:300

dilution overnight at 4°C, following avidin and biotin blocking. After AEC staining was

complete, slides were counterstained with a 1:2 dilution of Gill No. 1 hematoxylin solution for

30 seconds and then rinsed with tap water for 2 minutes. Coverslips were mounted using aqueous

Page 53: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

40

ImmunoHistoMount (Sigma). A positive TUNEL control was included for each tumour by

treating a section with TACs-Nuclease to generate DNA breaks in every cell. A positive cleaved

caspase-3 control slide was included with paraffin-embedded jurkat cells treated with apoptosis-

inducing etopiside (Sigma).

2.9 Immunohistochemistry Imaging and Quantitation

For all IHC analyses, digital images of slides were created using a Hamamatsu

NanoZoomer 2.0-RS Digital Slide Scanner (Meyer Instruments, Houston, TX, USA). Ki-67

Antigen (Ki-67) and proliferative cell nuclear antigen (PCNA) nuclear labeling index (LI) was

determined using the ImmunoRatio quantitative image analysis program (Isola J. and Tuominen

V., University of Tampere, Tampere, Finland) [101]. Each image was analyzed for the percent of

DAB-stained nuclei out of the total (DAB- and hematoxylin-stained) nuclei. Three randomly

selected 20× power field images per tumour, with areas of necrosis excluded, were analysed and

an average PCNA and Ki-67 nuclear LI was calculated per tumour. Data were compiled to create

an average PCNA and Ki-67 nuclear LI for SKOV3-M and SKOV3-Ve tumours.

Necrotic tumour area was determined by annotating hematoxylin stained whole sections

of tumours using the Hamamatsu NanoZoomer Digital Pathology viewer (Meyer Instruments).

Areas of necrotic tissue were outlined using a freehand area annotation and necrotic area was

divided by total tumour area to determine the percentage of necrotic tumour area based on cell

and tissue morphology. Necrosis morphology and necrotic tumour area annotations were verified

by gynecologic oncology pathologist Dr. Blaise Clarke of the University Health Network

(Toronto, ON, Canada). Double TUNEL/cleaved caspase-3 staining was quantitated by counting

the number of TUNEL and cleaved caspase-3 positive cells per mm2 necrotic and non-necrotic

area of each tumour at 40× power field. Analysis was blinded as treatment conditions were

unknown during quantitation.

MVD (number of microvessels (MV) per 1 mm2) was visualized by staining all tumours

for the endothelial cell markers CD34 and CD31, and quantitated using the methodology

described by the International Consensus in 2002 [102]. MV were counted, at 40× power field, in

the three most vascularized areas (0.74 mm2 per field) of the tumour. MV were defined as any

distinct CD34+, or CD31

+, cell or cell cluster and the presence of a lumen was not required.

Page 54: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

41

Large vessels with thick muscular walls and areas of necrosis were excluded from the count.

CD31 images were analyzed blindly as treatment condition for each tumour was unknown while

counting. An average CD34+ and CD31

+ MVD was calculated for each tumour and data were

compiled to create an average CD34+ and CD31

+ MVD for SKOV3-M and SKOV3-Ve tumours.

CD34+ and CD31

+ MVD counts were verified by gynecologic oncology pathologist Dr. Blaise

Clarke of the University Health Network (Toronto, ON, Canada).

SOD2 total positive intensity was analysed using the Positive Pixel Count v9 Algorithm

included with ImageScope 11.0 software (Aperio Technologies, Inc., Vista, CA, USA). This

algorithm counts the number of pixels within three intensity ranges (weak, positive, and strong)

to create a total intensity. Three non-necrotic areas of each SKOV3-M and SKOV3-Ve tumour

were analyzed and data were compiled to determine an average total SOD2 positive intensity for

each treatment group.

2.10 SOD2 Activity Assay

SKOV3-M and SKOV3-Ve3 cells were seeded at 80,000 cells per well in a 24-well plate

and treated with 1 µM CdCl2 24 hours post-seeding to induce Veph1 expression (3 wells were

seeded per cell line). Twenty-four hours after treatment, cells were harvested by scraping and

suspended in media on ice. Cells were then centrifuged for 10 minutes, at 4°C, at 1,500 g and the

pellet of cells was mechanically homogenized in a cold solution of 20 mM HEPES buffer pH 7.2

(containing 1mM ethylene glycol tetraacetic acid (EGTA; Sigma), 210 mM mannitol (BDH

Chemicals), and 70 mM sucrose (Mallinckrodt Chemical, St. Louis, MO, USA)). Homogenized

cells were centrifuged for 5 minutes, at 4°C, at 1,500 g and the supernatant was collected and

seeded into a 96-well plate with each replicate seeded into 3 wells. Total protein concentration

was determined using a NanoDrop 1000 spectrophotometer (Thermo Fisher Scientific). SOD2

activity was determined using the recommended protocol for the Superoxide Dismutase Assay

Kit (Cayman Chemical Company, Ann Arbor, MI, USA). Potassium cyanide (3 mM; Acros

Organics, NJ, USA) was added to all wells to inhibit the activities of both Cu/Zn-SOD (SOD1)

and extracellular SOD (SOD3), resulting in the detection of only mitochondrial SOD2 activity.

Formazan dye accumulation, indicating superoxide radical dismutation, was quantified by

measuring absorbance at 450 nm with a spectrophotometer. The calculated average of each

standard was used to create a linearized SOD standard rate. The equation of the standard curve

Page 55: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

42

was used to determine SOD2 activity (U/ml) which was normalized to total protein

concentration. This experiment was independently repeated three times.

2.11 Statistical Analyses

Student’s t-test (p<0.05) was used to analyze IHC and non-necrotic TUNEL

quantitations, average number of days to tumour formation, necrotic tumour area, and in vitro

Flag induction data. One sample t-test to compare the ratio of SKOV3-Ve3/SKOV3-M to a

theoretical mean of 1.0 (p<0.05) was used for SOD2 in vitro protein and activity data. Fisher’s

Exact test (p<0.05) was used to examine percentage tumour formation and tumour liquid

accumulation data. Two-way analysis of variance (ANOVA), followed by a Student-Newman-

Keuls (SNK) post-hoc test, was used to analyze necrotic area TUNEL and cleaved caspase-3

positive staining (p<0.05).

A Kaplan-Meier survival plot was created using GraphPad Prism version 5 (GraphPad

Software Inc., La Jolla, CA, USA) and analyzed using a Gehan-Breslow-Wilcoxon Test

(p<0.05). A growth rate plot indicating tumour volume (mm3) at each time point, with

exponential lines of best-fit, was created using GraphPad Prism version 5 (GraphPad Software

Inc.). Analysis of covariance was used to compare the linear regression lines generated for

SKOV3-M and SKOV3-Ve tumour growth as a function of time (p<0.05).

Page 56: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

43

2.3. Chapte

RESULTS

3.1 Veph1 Expression in Murine Embryos and Adult Tissues

3.1.1 Veph1A and Veph1B mRNA expression in murine whole embryos and adult tissues

To determine expression of Veph1 during development and its tissue distribution, the

presence of transcripts in whole embryos and adult tissues was examined by RT-PCR. Two

mRNA variants of Veph1 have been described in mice, one encoding the full-length protein

(Veph1A) and another (Veph1B) predicted to encode the C-terminal portion of Veph1A, including

the PH domain. Total RNA was extracted from murine embryos (E9.5-18.5) and various adult

tissues and RT-PCR was performed using primers specific to each isoform. Veph1A transcripts

were detected from embryonic day 10.5 to E18.5, whereas Veph1B expression was not detected

beyond E17.5 (Figure 5). Veph1A RNA was expressed in the following murine adult

reproductive tissues: ovary (detected in 1 of 3 independent tissue sets; not shown), oviduct,

epididymis, and testis, but was not detected in the uterus, ovarian fat pad, prostate, and seminal

vesicle. Veph1B expression was not detected in any adult reproductive tissue examined (Figure

6A). Veph1A RNA was expressed in the following murine adult non-reproductive tissues: liver,

lung, brain, kidney, and eye but was not detected in the heart, spleen, adrenal gland, or skeletal

muscle. Veph1B expression was not detected in any adult non-reproductive tissue examined

(Figure 6B). PCR amplicons were verified to be either Veph1A or Veph1B by automated

sequencing.

3.1.2 Veph1A protein expression and distribution in murine adult tissues

To determine the expression and localization of Veph1A protein, murine adult tissues

were examined by Western blot analysis and IHC staining. Currently, there are four

commercially available antibodies raised against human Veph1; however, only three of these

target murine Veph1A epitopes with greater than 90% sequence identity (Figure 7).

Optimization experiments indicated that the Santa Cruz rabbit polyclonal (SC-Rab-PC) and

Santa Cruz goat polyclonal (SC-Goat-PC) antibodies did not detect Veph1 protein in human

ovarian cancer HEY cells by Western blot analysis. HEY cells express high levels of endogenous

Chapter 3

Page 57: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

44

Figure 5: Veph1A and Veph1B transcript expression in murine whole embryos. Total RNA

was extracted from three independent sets of murine whole embryos and RT-PCR was

performed using primers specific to Veph1A, Veph1B, and ACTB (β-actin). Amplicons were

resolved on 2% agarose gels and shown to consist of a single product corresponding to the

expected sizes of the target sequences. A representative of 3 gels are shown. No RT control was

created by omission of reverse transcriptase during E18.5 cDNA synthesis. PCR control was

created by substituting double distilled water in place of cDNA during PCR.

Page 58: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

45

Figure 6: Veph1A and Veph1B transcript expression in murine adult reproductive and non-

reproductive tissues. Total RNA was extracted from three independent sets of murine adult

reproductive (A) and two to four independent sets (see Table #) of non-reproductive (B) tissues.

RT-PCR was then performed using primers specific to Veph1A, Veph1B, and ACTB. Amplicons

were resolved on 2% agarose gels and shown to consist of a single product corresponding to the

expected sizes of the target sequences. A representative of 3 gels are shown. No RT control was

created by omission of reverse transcriptase during testes (A) and kidney (B) cDNA synthesis.

PCR control was created by substituting double distilled water in place of cDNA during PCR.

Representative images are shown; positive ovary band not shown. Total RNA extracted from

E15.5 whole embryos was used as a positive control.

A

B

Page 59: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

46

Figure 7: Homology between murine and human Veph1 protein sequences and summary of

epitopes of the human Veph1 protein targeted by commercially available antibodies.

Protein sequence of human full-length Veph1 (isoform 1) aligned with murine Veph1A (Veph1B

start site indicated at amino acid 580). Sequence alignment was performed using the protein

Basic Local Alignment Search Tool from National Center for Biotechnology Information. Non-

identical amino acid residues between murine and human Veph1 are indicated in red text. PH

domain of human and murine Veph1 proteins is indicated in light blue text. Commercially

available antibodies are listed with manufacturer name, source (Rabbit-polyclonal, Rab-PC;

Goat-polyclonal, Goat-PC), identity of targeting antigen with murine Veph1 protein sequence,

and optimization details for both Western blot and IHC analyses. HEY cell lysates were used as

a positive control for Western blot analyses.

Page 60: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

47

Page 61: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

48

Veph1 (Shathasivam et al., in preparation; data not shown). Further, IHC optimization with the

SC-Rab-PC antibody indicated staining in the IgG control at the concentration required to detect

staining and the SC-Goat-PC antibody did not stain any tissue. The Sigma rabbit polyclonal

(Sigma-Rab-PC) antibody detected a band consistent with Veph1 in HEY lysates and staining

was evident in select tissue sections at concentrations that did not result in staining by normal

IgG. Therefore the Sigma-Rab-PC antibody was selected for use throughout this study.

IHC analyses revealed positive staining in select murine adult female (Figure 8) and

male (Figure 9) tissues with localization in oocytes and corpora lutea of the ovary, B

spermatogonia and late spermatids of the testis, stroma of the prostate, and the epithelium of the

oviduct, uterus, prostate, epididymis, and seminal vesicle. Positive Veph1A staining was also

identified in the ventricular zone of the brain, myofibril striations of skeletal muscle, select cells

of the spleen, bronchial epithelium of the lung, and epithelia of proximal and distal kidney

tubules. Lung, brain, and kidney staining are depicted in Figure 10 and heart, liver, spleen, and

skeletal muscle staining are shown in Figure 11. Reproductive tissue Veph1A RNA expression

and protein localization is summarized in Table 4 and non-reproductive is summarized in Table

5.

To analyze the specificity of the Sigma-Rab-PC antibody, Western blot analysis was

performed on two independent sets of murine adult tissues. An expected 95 kDa band in positive

control HEY human ovarian cancer cell lysates was observed; however, additional bands of

various molecular weights were detected in all murine tissues (Reproductive tissue Figure 12;

Non-reproductive tissue Figure 13).

3.2 Veph1 Expression Effect on Ovarian Tumour Progression

3.2.1 Veph1 effect on tumour formation

To determine if Veph1 affects tumour formation, nude mice were injected

subcutaneously with SKOV3 cells stably transfected with either a metallothionein-inducible

vector containing Flag-tagged Veph1 or an empty vector. Tumour formation was monitored

every 24-48 hours. Veph1 expression did not have a significant effect on the formation of end-

Page 62: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

49

Figure 8: Veph1A localization in murine adult female reproductive tissues. Veph1A

localization in the ovary (A; 10X), oviduct (C; 20X), and uterus (E; 20X) revealed by IHC

staining with Sigma-Rab-PC antibody targeting Veph1A. Brown staining indicates Veph1

expression. IgG control sections were incubated with an equivalent amount of rabbit IgG as

primary antibody (B 10X, D 10X, and F 20X). Scale bars are 50-100 µm as indicated.

100 µm

100 µm 50 µm

50 µm

A

C D

E

100 µm

B

50 µm

F

Page 63: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

50

Figure 9: Veph1A localization in murine adult male reproductive tissues. Veph1A

localization in the prostate (A; 20X), seminal vesicle (C; 20X), epididymis (E; 100X), and testis

(G; 100X) revealed by IHC staining with Sigma-Rab-PC antibody targeting Veph1A. Brown

staining indicates Veph1 expression. IgG control sections were incubated with an equivalent

amount of rabbit IgG as primary antibody (B 20X, D 10X, F 10X, and H 100X). Scale bars are

10-100 µm as indicated.

Page 64: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

51

50 µm

50 µm

100 µm

100 µm 10 µm

A

C D

E F

G H

50 µm

B

Page 65: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

52

Figure 10: Veph1A localization in murine adult non-reproductive tissues. Veph1A

localization in the lung (A; 20X), brain (C; 20X), proximal kidney (E; 20X), and distal kidney

(G; 20X) revealed by IHC staining with Sigma-Rab-PC antibody targeting Veph1A. Brown

staining indicates Veph1 expression. IgG control sections were incubated with an equivalent

amount of rabbit IgG as primary antibody (B 10X, D 10X, F 10X, and H 10X). Scale bars are

50-100 µm as indicated.

Page 66: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

53

50 µm 100 µm

50 µm 100 µm

100 µm

A B

C D

E F

G H

Page 67: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

54

Figure 11: Veph1A localization in murine adult non-reproductive tissues. Veph1A

localization in the spleen (A; 40X) and skeletal muscle (C; 20X; IgG staining identified)

revealed by IHC staining with Sigma-Rab-PC antibody targeting Veph1A. Brown staining

indicates Veph1 expression. Veph1A expression was negative in the heart (E; 20X) and liver (G;

20X). IgG control sections were incubated with an equivalent amount of rabbit IgG as primary

antibody (B, D, F, and H; 10X). Scale bars are 20-100 µm as indicated.

Page 68: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

55

50 µm 100 µm

50 µm 100 µm

A B

C D

E F

G H

Page 69: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

56

Table 4: Summary of Veph1A mRNA and protein expression, and protein localization, in

murine adult male and female reproductive tissues.

Tissue RT-PCR

(Veph1A)*

IHC

(Veph1A)

IHC Veph1A Localization

Ovary 1/3 Positive -Oocytes

-Corpus luteum

Oviduct 2/3 Positive -Fallopian tube epithelium

Uterus 0/3 Low to

undetectable

-Endometrial + uterine glandular

epithelium

Prostate 0/3 Positive -Prostate epithelium

-Stroma

Seminal

Vesicle

0/3 Positive -Select seminal vesicle epithelium

Epididymis 3/3 Positive -Epididymal epithelium

Testes 3/3 Positive -B spermatogonia

-Late spermatids

Veph1A transcripts were detected by RT-PCR, and protein by IHC staining, in murine adult male

and female reproductive tissues. * column with RT-PCR data indicates number of independent

tissue sets positive for Veph1A from the total number of tissue sets examined.

Page 70: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

57

Table 5: Summary of Veph1A mRNA and protein expression, and protein localization in

murine adult non-reproductive tissues.

Tissue RT-PCR

(Veph1A)*

IHC

(Veph1A)

IHC Veph1A Localization

Liver 2/4 Negative Negative

Lung 3/3 Positive -Bronchial epithelium

Brain 2/4 Positive -Ventricular zone

Spleen 0/3 Positive -Select cells

Adrenal

Gland

0/3 N/A N/A

Kidney 3/3 Positive -Proximal + distal tubule

epithelium

Eye 2/2 N/A N/A

Skeletal

Muscle

0/2 Positive -Myofibril striations

Veph1A transcripts were detected by RT-PCR, and protein by IHC staining, in murine adult non-

reproductive tissues. * column with RT-PCR data indicates number of independent tissue sets

positive for Veph1A from the total number of tissue sets examined. Tissues that were not

analyzed via immunohistochemistry are indicated (N/A).

Page 71: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

58

Figure 12: Veph1A protein expression in murine adult male and female reproductive

tissues, as determined by Western blot analyses. Total protein was extracted from two

independent sets of selected murine adult male and female reproductive tissues and examined by

Western blot analysis using Sigma-Rab-PC antibody to detect Veph1A. Protein lysates (40 µg of

tissue lysate; 8 µg of HEY positive control lysate) were resolved by 10% SDS-polyacrylamide

gel electrophoresis (PAGE), transferred to a PVDF-membrane, and probed with Veph1 antibody

(1:100). Tissues that were positive for Veph1A RNA, detected by RT-PCR, are indicated (*).

GAPDH (1:2000) was used as a loading control.

Page 72: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

59

Page 73: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

60

Figure 13: Veph1A protein expression in murine adult non-reproductive tissues, as

determined by Western blot analyses. Total protein was extracted from two independent sets

of selected murine adult non-reproductive tissues and examined by Western blot analysis using

Sigma-Rab-PC antibody to detect Veph1A. Protein lysates (40 µg of tissue lysate; 8 µg of HEY

positive control lysate) were resolved by 10% SDS-PAGE, transferred to a PVDF-membrane,

and probed with Veph1 antibody (1:100). Tissues that were positive for Veph1A RNA, detected

by RT-PCR, are indicated (*). GAPDH (1:2000) was used as a loading control.

Page 74: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

61

Page 75: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

62

point volume tumours (520 mm3), growths of less than 520 mm

3, or non-respondent tumours

between groups (p=1.00; Figure 14).

To verify that supplementation of 25 mM ZnSO4 in the drinking water (provided to all

mice) resulted in Flag-tagged Veph1 expression in SKOV3-Ve tumours, Western blot analysis

was conducted on all tumours obtained. As expected, Flag expression was detected in all

SKOV3-Ve tumours and was absent in SKOV3-M tumours (Figure 15).

3.2.2 Veph1 effect on tumour growth rate and survival

To determine if Veph1 expression altered in vivo growth rate of SKOV3 cells, tumour

dimensions were determined every 24-48 hours via digital caliper measurements. SKOV3-M

tumours advanced to endpoint volume (520 mm3) tumours more rapidly than SKOV3-Ve

tumours (31.6 days vs. 45.8 days; p<0.001; Figure 16A). Kaplan-Meier survival curves for the

two groups of mice indicated a significant difference due to SKOV3 Veph1 expression with mice

injected with SKOV3-Ve cells having greater survival (death resulting from euthanasia due to

end-point volume tumour size or 63 days post-injection; p=0.0097; Figure 16B). The average

growth rate of SKOV3-M tumours (n=9; 27.46 +/- 1.792 mm3/time) was significantly greater

than SKOV3-Ve (n=8; 17.50 +/- 0.822 mm3/time) at each time point. This was determined by

analysis of covariance of linear regression lines depicting tumour growth as a function of time

(p<0.001; Figure 16C). Plotting the growth pattern of each tumour revealed a similar, yet

delayed, exponential growth pattern for SKOV3-M and SKOV3-Ve tumours (Figure 16D).

3.2.3 Veph1 effect on proliferative index

To determine whether the slower growth rate of SKOV3-Ve tumours was due to a

Veph1-dependent reduction in proliferative index, all tumours were immunohistochemically

stained for the proliferative markers PCNA and Ki-67. IHC analysis revealed no significant

difference in cycling cells (Ki-67 labeling index) between SKOV3-M (15.2 +/- 0.8%) and

SKOV3-Ve (13.9 +/- 1.5%) tumours (p=0.448; Figure 17). Similarly, there was no significant

difference in S-phase cells (nuclear PCNA labeling index) between SKOV3-M (34.8 +/- 2.6%)

and SKOV3-Ve (40.3 +/- 2.1%) tumours (p=0.126; Figure 18).

Page 76: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

63

Figure 14: SKOV3-M and SKOV3-Ve tumour formation. Graph demonstrating percentage of

mice that formed an endpoint volume tumour (>520 mm3; red) and percentage of mice that

formed any tumour growth (ie. not including non-respondent mice; blue) from a total of 12 mice

injected initially in the SKOV3-M and SKOV3-Ve treatment groups. Tumour formation data

were analyzed using a Fisher’s Exact test.

Page 77: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

64

Figure 15: ZnSO4-supplemented drinking water induction of Flag-Veph1. Total protein

lysates were extracted from all excised tumours and examined by Western blot analysis to detect

Flag-Veph1 expression using a Flag antibody. Protein lysates (20 µg) were resolved by 8% SDS-

PAGE, transferred to a PVDF-membrane, and probed with mouse monoclonal Flag antibody

(1:500). The membranes were then probed for GAPDH (1:2000) as a loading control. Graphs

demonstrate a quantitative presentation of Flag-Veph1 protein levels observed in tumours,

normalized to GAPDH levels. Each bar label represents mouse identification numbers.

Page 78: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

65

Page 79: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

66

Figure 16: SKOV3-M and SKOV3-Ve tumour growth rate and survival. A) Number of days

to formation of endpoint volume (>520 mm3) SKOV3-M (n=9) and SKOV3-Ve (n=8) tumours.

Data are shown as mean +/- standard error of mean (SEM) and were analyzed using Student’s t-

test. * indicates significant difference relative to SKOV3-M tumours (p<0.05). B) Kaplan-Meier

survival curve indicating percentage of surviving SKOV3-M (n=11) and SKOV3-Ve (n=10)

mice as a function of time (ie. tumours have not reached endpoint of >520 mm3) at each time

point. Data were analyzed using a Gehan-Breslow-Wilcoxon Test (p<0.05). C) Graph

demonstrating the exponential growth rate of SKOV3-M (n=9) and SKOV3-Ve (n=8) tumours

that reached endpoint volume; tumour volumes at each time point are indicated. Analysis of

covariance was used to compare the linear regression lines generated for SKOV3-M and

SKOV3-Ve tumour growth as a function of time (p<0.05). D) Graph demonstrating the

individual growth rate of all SKOV3-M (n=11) and SKOV3-Ve (n=10) tumours; tumour

volumes at each time point are indicated.

Page 80: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

67

A

B

C

D

Page 81: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

68

Figure 17: SKOV3-M and SKOV3-Ve tumour nuclear Ki-67 proliferative labeling index.

A) Nuclear Ki-67 proliferative labeling index (number of Ki67 stained nuclei compared to total

number of nuclei) for SKOV3-M (n=10) and SKOV3-Ve (n=10) tumours. Data are shown as

mean +/- SEM and were analyzed using a Student’s t-test (p<0.05). Representative SKOV3-M

(B) and SKOV3-Ve (C) mouse monoclonal Ki-67 (1:200) stained tumour images are presented.

IgG control section was incubated with normal mouse IgG rather than primary antibody at an

equivalent amount as primary antibody (D).

A

B

C

D

Page 82: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

69

A

B

A

Figure 18: SKOV3-M and SKOV3-Ve tumour nuclear PCNA proliferative labeling index.

A) Nuclear PCNA proliferative labeling index (number of PCNA stained nuclei compared to

total number of nuclei) for SKOV3-M (n=11) and SKOV3-Ve (n=10) tumours. Data are shown

as mean +/- SEM and were analyzed using a Student’s t-test (p<0.05). Representative SKOV3-M

(B) and SKOV3-Ve (C) mouse monoclonal PCNA (1:2000) stained tumour images are

presented. IgG control section was incubated with normal mouse IgG rather than primary

antibody at an equivalent amount as primary antibody (D).

C

D

Page 83: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

70

3.2.4 Veph1 effect on necrotic tumour area

To examine morphology, all tumours were stained for hematoxylin and eosin.

Microscopic observation of these images revealed necrotic areas in all tumours. Analysis of

tumour necrotic area, based on morphology, revealed a 55.0% greater average area of necrosis in

SKOV3-Ve tumours (24.3 +/- 5.4%), as compared to SKOV3-M tumours (11.0 +/- 2.7%;

p=0.034; Figure 19A). Necrosis morphology was verified by gynecologic oncology pathologist

Dr. Blaise Clarke of the University Health Network. Upon tumour excision it was noted that

60% of SKOV3-Ve tumours collected were fluid filled, often associated with necrosis, while

none of the SKOV3-M tumours contained fluid (p=0.004; Figure 19B).

To verify whether areas of apparent necrosis identified actually reflected necrosis or

apoptosis, double staining for TUNEL, used to detect DNA fragmentation in dying cells, and

cleaved caspase-3, which is present in cells undergoing apoptosis, was conducted on all tumours.

Analysis of TUNEL staining revealed a 64.9% greater average number of TUNEL positive cells

per necrotic area in Veph1 expressing (390.30 +/- 84.40 TUNEL+

cells/mm2 necrotic area), as

compared to non-expressing tumours (136.92 +/- 46.16 TUNEL+ cells/mm

2 necrotic area;

p<0.001). IHC analysis revealed no difference in cleaved caspase-3 staining between SKOV3-M

(20.07 +/- 8.89 cleaved caspase-3+ cells/mm

2 necrotic area) and SKOV3-Ve (20.73 +/- 9.12

cleaved caspase-3+ cells/mm

2 necrotic area) tumours. Fewer cleaved caspase-3

+ cells, as

compared to TUNEL+ cells, were identified in necrotic areas of SKOV3-M (85.68% reduction)

and SKOV3-M (99.95% reduction) tumours (p<0.001; Figure 20A). A greater (46.6%) average

number of TUNEL positive cells per non-necrotic area was also identified in SKOV3-Ve

tumours (6.82 +/- 2.49 TUNEL+ cells/mm

2 non-necrotic area), as compared to SKOV3-M

tumours (0.73 +/- 0.19 TUNEL+ cells/mm

2 non-necrotic area) (p=0.04; Figure 20D). No cleaved

caspase-3 staining was identified in non-necrotic tumour areas.

3.2.5 Veph1 effect on microvessel density

To determine if there was an effect of Veph1 expression on MVD, tumours were

immunohistochemically stained for the endothelial cell markers CD31 and CD34. Quantitation

of CD34 staining revealed a 50.6% decrease in MVD in SKOV3-Ve tumours (27.7 +/-5.6

MV/mm2), as compared to SKOV3-M tumours (56.1 +/- 6.5 MV/mm

2; p=0.004; Figure 21).

Page 84: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

71

Figure 19: SKOV3-M and SKOV3-Ve tumour necrotic area and liquid accumulation. A)

Average necrotic tumour area (percent of necrotic area as compared to total tumour area;

determined by annotation based on morphology) for SKOV3-M (n=11) and SKOV3-Ve (n=10)

tumours. Data are shown as mean +/- SEM and were analyzed by Student’s t-test. B) Percentage

of tumours filled with liquid, that is commonly associated with necrosis, for SKOV3-M (n=11)

and SKOV3-Ve (n=10) treatment groups. Data were analyzed by Fisher’s Exact test (p<0.05). *

indicates significant difference relative to SKOV3-M tumours (p<0.05). Representative SKOV3-

M (C) and SKOV3-Ve (D) tumours stained with hematoxylin and eosin reflect necrotic areas

(indicated by black arrows).

D C

A B

Page 85: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

72

Figure 20: SKOV3-M and SKOV3-Ve tumour double TUNEL and cleaved caspase-3

staining. A) Average number of TUNEL and cleaved caspase-3 positive cells per necrotic

tumour area (mm2) for SKOV3-M (n=11) and SKOV3-Ve (n=8) tumours. Bars with different

letters are statistically significant from one another as determined by ANOVA followed by a

SNK post-hoc test (p<0.05). Representative SKOV3-M (B) and SKOV3-Ve (C) necrotic tumour

areas stained for TUNEL and cleaved caspase-3 are shown. Green arrows indicate several

cleaved caspase-3 stained cells. D) Graph demonstrating average number of TUNEL positive

cells per non-necrotic tumour area (mm2) for SKOV3-M (n=11) and SKOV3-Ve (n=8) tumours.

Data were analyzed by Student’s t-test. * indicates significant difference relative to SKOV3-M

tumours (p<0.05). Representative SKOV3-M (E) and SKOV3-Ve (F) non-necrotic tumour areas

stained for TUNEL and cleaved caspase-3 are shown. Red arrows indicate several TUNEL

positive tumour cells. Data are shown as mean +/- SEM.

Page 86: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

73

A

D

B

C

E

F

Page 87: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

74

A

Figure 21: SKOV3-M and SKOV3-Ve tumour microvessel density as determined by CD34

staining. A) Average MVD (number of CD34+ microvessels per mm

2 tumour area) for SKOV3-

M (n=11) and SKOV3-Ve (n=10) tumours. Data are shown as mean +/- SEM and were analyzed

by Student’s t-test. * indicates significant difference relative to SKOV3-M tumours (p<0.05).

Representative SKOV3-M (B) and SKOV3-Ve (C) rabbit monoclonal CD34 (1:250) stained

tumour images are presented with red arrows indicating CD34+ MV. IgG control section was

incubated with normal rabbit IgG rather than primary antibody at an equivalent amount as

primary antibody (D).

B

C

D

Page 88: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

75

Quantitation of CD31 staining revealed a similar 35.2% decrease in MVD in SKOV3-Ve

tumours (31.4 +/- 3.6 MV/mm2), as compared to SKOV3-M tumours (48.4 +/- 6.7 MV/mm

2)

(p=0.041; Figure 22).

3.2.6 Veph1 effect on SOD2 protein expression and activity

The effect of Veph1 on SOD2 protein expression was examined in vitro and in vivo as

SOD2 has a potential function in promoting angiogenesis [71]. A significant 86.4% reduction in

SOD2 protein expression was identified in SKOV3-Ve3 cells, as compared to SKOV3-M (three

independent repeats; p=0.0046; Figure 23A). A reduction in SOD2 protein levels was also seen

in the SKOV3-Ve4 clone, as compared to SKOV3-M cells (Figure 23B). A 75.5% reduction in

SOD2 activity, as compared to SKOV3-M cells (three independent repeats; p=0.0008; Figure

23C), was identified in SKOV3-Ve3 cells which correlates with the decrease in SOD2 protein

levels. Confirmation of the induction of Flag-Veph1 expression following 24 and 48 hour CdCl2

treatment in SKOV3-Ve3 cells was evident by Western blot detection using an anti-Flag

antibody (24 hours, p=0.034; 48 hours, p=0.026; Figure 23D).

To determine whether this effect of Veph1 on decreasing the protein expression of SOD2

was maintained in vivo, all xenograft tumours were examined via Western blot and IHC

analyses. Western blot analysis did not reveal a significant difference in SOD2 expression

between SKOV3-M and SKOV3-Ve tumours (Figure 24). Tumour sections were stained for

SOD2 to analyze total staining intensity in SKOV3 tumour cells. Quantitation of SOD2 staining

also did not reveal a significant difference in protein expression between SKOV3-M (1.64x107

+/- 2.67x106 total staining intensity) and SKOV3-Ve tumours (1.93x10

7 +/- 2.71x10

6 total

staining intensity) (p=0.442; Figure 25).

Page 89: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

76

Figure 22: SKOV3-M and SKOV3-Ve tumour microvessel density as determined by CD31

staining. A) Average MVD (number of CD31+ microvessels per mm

2 tumour area) for SKOV3-

M (n=11) and SKOV3-Ve (n=10) tumours. Data are shown as mean +/- SEM and were analyzed

by Student’s t-test. * indicates significant difference relative to SKOV3-M tumours (p<0.05).

Representative SKOV3-M (B) and SKOV3-Ve (C) rabbit monoclonal CD31 (1:50) stained

tumour images are presented with red arrows indicating CD31+ MV (IHC staining was

performed by Pathology Services at The Toronto Centre for Phenogenomics).

B

A

C

Page 90: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

77

Figure 23: SKOV3-M and SKOV3-Ve SOD2 expression and activity in vitro. A) SKOV3-M

and SKOV3-Ve3 cells were grown in the presence or absence of 1 µM CdCl2 for 24 hours and

total protein was extracted and examined by Western blot for SOD2 expression. Data were

analyzed using a one sample t-test to compare the ratio of SKOV3-Ve3/SKOV3-M SOD2

protein levels to a theoretical mean of 1.0 (three independent experiments). This experiment was

repeated in the SKOV3-Ve4 clone once (24 and 48 hour treatments with 1 µM CdCl2) to confirm

the effect of Veph1 on SOD2 expression (the membranes were probed for Flag to confirm

CdCl2-induction of Flag-Veph1; B). C) SKOV3-M and SKOV3-Ve3 cells were grown in the

presence or absence of 1 µM CdCl2 for 24 hours and SOD2 activity was determined (three

replicates per experiment; three independent experiments). The equation of the standard curve

was used to determine SOD2 activity (U/ml) and SOD2 activity was normalized to total protein

concentration. Data were analyzed using a one sample t-test to compare the ratio of SKOV3-

Ve3/SKOV3-M SOD2 activity levels to a theoretical mean of 1.0. D) SKOV3-M and SKOV3-

Ve3 cells were grown in the presence or absence of 1 µM CdCl2 for 24 and 48 hours. Total

protein was extracted and examined by Western blot for Flag expression. Data were analyzed by

Student’s t-test (three independent experiments). Protein lysates (8 µg) were resolved by 12%

SDS-PAGE in A and B, and 8% SDS-PAGE in D. The membranes in A, B, and D were probed

for GAPDH (1:2000) as a loading control. Graphs demonstrate a quantitative presentation of

protein levels, normalized to GAPDH levels. Graphs in A, C, and D present as mean +/- SEM of

three independent experiments. * indicates a statistically significant difference relative to

SKOV3-M cells (p<0.05).

Page 91: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

78

A B

C

D

Page 92: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

79

Figure 24: SOD2 protein expression in SKOV3-M and SKOV3-Ve tumours, as determined

by Western blot analysis. Total protein lysates were extracted from all excised tumours and

examined by Western blot analysis to detect SOD2 expression. Protein lysates (20 µg) were

resolved by 12% SDS-PAGE, transferred to a PVDF-membrane, and probed with rabbit

polyclonal SOD2 antibody (1:5000). The membranes were also probed for GAPDH (1:2000) as a

loading control. Graphs demonstrate a quantitative presentation of SOD2 protein levels observed

in tumours that were normalized to GAPDH levels. Each bar label represents mouse

identification numbers.

Page 93: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

80

Page 94: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

81

A

Figure 25: SOD2 protein expression in SKOV3-M and SKOV3-Ve tumours, as determined

by immunohistochemical analysis. A) Average total intensity of SOD2 staining for SKOV3-M

(n=10) and SKOV3-Ve (n=10) tumours. Data are shown as mean +/- SEM and were analyzed by

Student’s t-test. Representative SKOV3-M (B) and SKOV3-Ve (C) rabbit monoclonal SOD2

(1:250) stained tumour images. IgG control section was incubated with normal rabbit IgG rather

than primary antibody at an equivalent amount as primary antibody (D).

B

C

D

Page 95: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

82

2.4. Chapte

DISCUSSION

4.1 Summary of Findings

Veph1 may alter the expression of genes that are partially regulated by the FoxO, TGF-β,

and Hippo signalling pathways that have both ovarian tumour suppressing and promoting

functions (Shathasivam et al., in preparation). Signalling modulation by Veph1 is supported by

studies conducted in human and Drosophila cells [8-11]. A role for Veph1 in EOC is strongly

implicated by genomic analyses that indicate an increase in VEPH1 gene copy number in a

subset of ovarian primary tumours and cell lines [4, 5]. Furthermore, Veph1 expression

correlates with an invasive and motile phenotype in ovarian cancer cell lines (Shathasivam et al.,

in preparation) [6, 7]. Altogether these data indicate Veph1 may have both physiological and

tumourigenic roles.

In this thesis, expression of Veph1A/Veph1B transcripts and protein was investigated in

murine embryonic and adult tissues. Expression of Veph1 transcripts was more widespread than

initially reported [12]. Adult murine Veph1A protein localization appeared to be largely

restricted to epithelium; however, further studies with more specific antibodies are required.

Furthermore, the effect of Veph1 expression on SKOV3 primary tumour formation and

progression was established. Expression of Veph1 reduced the growth rate of tumours; however,

this was not due to an effect on proliferative capacity. A significant impairment in angiogenesis

was identified in SKOV3 cell xenograft tumours that expressed Veph1. The resultant low MVD

in SKOV3-Ve tumours may suggest a mechanism by which these tumours exhibited a significant

increase in necrotic tumour area associated with liquid accumulation. Lastly, Veph1 expression

resulted in a significant reduction in SOD2 protein and activity levels in vitro; however, these

findings were not replicated in vivo.

4.2 Widespread Physiological Expression of Veph1 in Mice

Veph1 has a critical function in the development of the zebrafish central nervous system

[12]. In addition, the Drosophila orthologue of Veph1, Melted, has been identified to have an

imperative role peripheral nervous system development [60, 61]. However, tissue expression

Chapter 4

Page 96: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

83

data are limited. Muto et al. [12] identified murine Veph1 mRNA in the developing embryonic

eye and brain, particularly in the ventricular zone. Adult murine Veph1 tissue expression is

maintained in the eye and kidney while other examined tissues appear to be negative for Veph1

transcript expression [12]. Importantly, murine Veph1 protein tissue expression and localization

has not been investigated.

Microarray data generated previously in our lab indicate that Veph1 protein expression in

SKOV3 human ovarian cancer cells alters the expression of genes that are partly regulated by the

FoxO, TGF-β, and Hippo signalling pathways (Shathasivam et al., in preparation). As these

signalling pathways have tumour suppressing and promoting roles, it was of interest to establish

the physiological expression of murine Veph1 transcript and protein during embryonic

development and adulthood. While Muto et al. [12] examined murine Veph1 mRNA expression

using primers that detect both transcript variants, my study used primers that distinguish between

Veph1A and Veph1B.

My studies reveal that Veph1A and Veph1B transcript expression is maintained

throughout embryogenesis, suggesting a potential role for both Veph1 isoforms during

embryonic development. Interestingly, Veph1B expression does not appear to be maintained in

adult tissues which may be indicative of an embryonic-specific role. This is of interest as our lab

has identified that a truncated protein consisting of amino acids 662-833 of full-length human

Veph1, which is similar to Veph1B (amino acids 580-833 of Veph1A), promotes TGF-β

signalling. This is in contrast to the full-length protein which inhibits TGF-β signalling. Further,

co-expression of this truncated peptide prevents the inhibition of TGF-β signalling by full-length

Veph1 (Shathasivam et al., in preparation). These data suggest that Veph1B may modulate TGF-

β signalling during embryonic development.

Morpholino-based knockdown of the zebrafish Veph1 gene (zVeph1) in fertilized eggs

results in embryos with defective central nervous system development, including specific

impairment in the formation of the midbrain–hindbrain boundary and otic vesicle formation [12].

Furthermore, mutation of melted in Drosophila results in abnormal morphology and loss of

peripheral neurons [61]. Due to possible species conservation, these data suggest that mutation of

murine Veph1 may also result in impaired brain development. Generation of Veph1 knockout

Page 97: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

84

mice has revealed no defects in embryonic or adult development [12]. However, this lack of

effect of Veph1 on phenotype may be due to the region of gene deletion. Mice were generated by

deleting exons encoding amino acids 576-710 thus preserving the PH domain in both Veph1A

and Veph1B and a majority of the N-terminus of Veph1A [12]. Expression of these regions

could potentially preserve the function of murine Veph1 isoforms and allow normal

development. Therefore, a knockout of full length Veph1 would be required to assess its impact

in mammalian brain development.

Veph1A was expressed in the adult kidney and eye in my studies, consistent with

previous reports [12]. These data indicate that adult kidney and eye expression reported by Muto

et al. [12] represented Veph1A isoform specifically. In contrast to the murine transcript data

previously reported, I identified widespread expression of Veph1A in adult tissues. This

discrepancy may be due to procedural differences. It is important to note that PCR products

identified in my study were sequence verified. As my data suggest adult murine tissue expression

of Veph1A is widespread, it was then imperative to establish whether identified transcripts are

translated.

There are four commercially available antibodies that target human Veph1, with three

having >90% epitope sequence identity with murine Veph1A. One of these antibodies was used

in this study as it revealed IHC staining and identified Veph1 in positive control HEY cells by

Western blot analysis; however, several unexpected immunoreactive bands of different

molecular weights were also identified in murine tissues. These bands may represent different

isoforms, non-specific bands due to poor antibody specificity, or degradation products.

There is evidence that supports the validity of the Sigma-Rab-PC antibody. In contrast to

several bands identified in mouse tissues by Western blot analyses, one band at the expected 95

kDa molecular weight for human Veph1 was identified in positive control HEY ovarian cancer

cells (data not shown). Furthermore, several immunoreactive bands were detected in the heart

and liver; however, no IHC staining was observed in these tissues suggesting that the Sigma-

Rab-PC antibody is not detecting these bands when used for immunohistochemical analyses.

Additionally, Veph1 mRNA has been previously identified in the ventricular zone of the murine

fetal brain [12] and this specific region stained positive for Veph1A in adult murine brain tissue

in my studies. Lastly, specificity of this antibody is supported by immunocytofluorescence

Page 98: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

85

detection of green fluorescent protein (GFP)-tagged Veph1 that is in agreement with Veph1

staining identified with the Sigma-Rab-PC antibody in SKOV3 cells (Shathasivam personal

communication). Staining identified in my studies will need to be verified upon the production of

more specific antibodies directed toward murine Veph1A. In situ hybridization to detect murine

Veph1 transcripts in adult tissues would also verify both mRNA expression and protein

localization data described in this study.

Positive Veph1A staining in adult tissues was localized to the epithelia of all tissues

examined, except for that of the ovary. Additional areas that were stained include specific

regions of the brain, corpus lutea and oocytes in the ovary, and B spermatogonia and late

spermatids in the testis. Melted and zVeph1, human Veph1 homologs, have imperative functions

in the development of the nervous system [12, 61]. My studies suggest that murine Veph1 may

have a function in the brain that is maintained during adulthood. In addition, Veph1A positive

staining of reproductive tissues suggests additional functions.

Veph1A positive staining was identified in oocytes and B spermatogonia which require

FoxO signalling for partial regulation of oogenesis and spermatogenesis, respectively. FoxO1

expression is required for spermatogonial stem cell maintenance as conditional FoxO1

knockdown in murine germ cells results in defective proliferative expansion of spermatogonia

[103]. FoxO1 signalling promotes stem cell maintenance by upregulating the transcription of

stem cell markers; however, by an unknown mechanism, FoxO1 becomes inactivated allowing

spermatogenic differentiation [103]. It is possible that Veph1 may be involved in stem cell

differentiation as its expression is induced upon the differentiation of human testis embryonal

carcinoma stem cells [12]. Veph1 may be involved in the mechanism by which FoxO is

inactivated during spermatogonial stem cell differentiation as Melted is reported to inhibit FoxO

activity [8] and Veph1 was specifically expressed in differentiated B spermatogonia in my study.

Veph1 may also be preventing FoxO signalling in oocytes.

FoxO signalling acts to prevent oocyte survival. The accumulation of active nuclear

FoxO3a results in apoptosis of rat oocytes that is accomplished by increased expression of pro-

apoptotic proteins- such as Bim and p27kip1 [104]. Although FoxO3a-induced oocyte apoptosis

is necessary to prevent excessive follicular activation, uncontrolled FoxO activation may result

in premature ovarian failure [104]. In my studies, Veph1A staining was identified in oocytes,

Page 99: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

86

suggesting a possible mechanism by which FoxO activity is inhibited and oocyte survival is

maintained.

Veph1 staining was identified most commonly in the epithelia of tissues examined. This

localization was predicted prior to this study as our lab has identified upregulation of epithelial

marker E-cadherin mRNA and protein in Veph1 expressing cells (Shathasivam et al., in

preparation). Veph1 was identified in FTE, however, not in OSE; E-cadherin is expressed in

normal FTE [105], neoplastic OSE, and inclusion cysts, but not in normal OSE [106].

Correlating localization data support the possible function of Veph1 in promoting E-cadherin

expression.

E-cadherin levels are reduced in metastatic ovarian tumour cells isolated from ascites as

compared to primary solid tumours from the same patient [107]. Further, increased E-cadherin

expression is associated with reduced invasive capability due to its prevention of EMT that is

necessary for ovarian cancer metastasis [107]. In addition to a function in preventing invasion, E-

cadherin expression has been shown to promote proliferation of SKOV3 cells which is prevented

by treating with an siRNA targeting the CDH1 gene [108]. These data suggest that Veph1-

dependent upregulation of E-cadherin may promote primary tumour growth while impairing

metastatic development. Establishing the effect of Veph1 on primary ovarian tumour growth was

the next objective of this thesis.

4.3 Veph1 Impedes SKOV3 Primary Tumour Growth

In this study, Veph1 unexpectedly reduced the growth rate of primary SKOV3 tumours in

vivo. This effect of Veph1 on tumour growth rate was not due to effects on the proliferative

capacity of cells; rather, it is likely due to a Veph1-induced impairment on angiogenesis that

causes a net promotion of tumour cell death and ultimate deficiency in overall growth.

Microarray data established in our lab indicate Veph1 affects the FoxO, TGF-β, and Hippo

signalling pathways. These pathways have been implicated in regulating angiogenesis and thus

there are several possible mechanisms by which Veph1 impairs angiogenesis.

The Drosophila Veph1 homolog Melted has been identified as an inhibitor of FoxO

activity [8]. The FoxO signalling pathway promotes SOD2 expression [65] which aids in the

production of H2O2 that regulates signalling pathways involved in promoting blood vessel

Page 100: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

87

formation [71]. H2O2 exposure can cause activation of the transcription factors HIF-1α, activator

protein 1 (AP-1), and v-ets erythroblastosis virus E26 oncogene homolog 1 (Ets-1), that

collectively cause the upregulation of genes encoding proteins involved in promoting

angiogenesis- such as VEGF, MMPs, and urokinase plasminogen activator (uPA) [71]. By

inhibiting FoxO-mediated SOD2 expression, Veph1 may reduce H2O2 production thereby

reducing MVD. I established that Veph1 causes a significant reduction in SOD2 protein and

activity levels in SKOV3 cells in vitro; however, this was not maintained in vivo. This loss of

Veph1-dependent SOD2 reduction may be due to differences between tissue culture conditions

and the in vivo tumour microenvironment. This inconsistency suggests that Veph1 may act via

another mechanism to reduce MVD. Alternatively, Veph1 may exert its effects on impeding

angiogenesis by its disruption of the TGF-β signalling pathway.

Our lab has established that Veph1 is a novel inhibitor of canonical TGF-β signalling

(Shathasivam et al., in preparation). TGF-β promotes angiogenesis in part by increasing the gene

expression of potent pro-angiogenic factor VEGF [40] and by promoting EndMT resulting in the

generation of cancer-associated fibroblasts that release angiogenesis-promoting factors including

TGF-β and VEGF [87]. Inhibition of TGF-β signalling decreases the expression of VEGF that

ultimately reduces CD31+ MV formation in human xenograft tumours [40]. Tissue microarray

analyses of 339 primary ovarian tumours revealed that increased VEGF protein expression was

significantly associated with poor survival [109]. Importantly, only 7% of tumours exhibited

high expression of VEGF suggesting that anti-angiogenic drugs may only be successful in a

subset of patients with ovarian cancer [109]. Analyses of microarray data conducted in our lab to

investigate the effect of Veph1 on gene expression of SKOV3 cells indicate that Veph1 causes a

significant decrease in VEGFA RNA levels (Shathasivam et al., in preparation). This effect has

not yet been validated but may suggest a possible mechanism by which Veph1 reduced MVD in

my studies.

Another possible mechanism by which Veph1 exerts its effects on angiogenesis is via its

potential impact on the Hippo signalling pathway. Melted is a negative regulator of Hippo

signalling [9-11] and our lab has identified that Veph1 alters the expression of several YAP-

regulated genes (Shathasivam et al., in preparation). The transcription factor YAP has been

described as an oncogene due to its role in increased cell proliferation, migration, survival, and

Page 101: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

88

chemoresistance [90]; however, a role in angiogenesis regulation has not been described.

Inhibition of LATS1 and LATS2 in ovarian cancer cells and primary tumours results in an

increased expression of YAP-regulated pro-angiogenic factor CTGF [92]. Our microarray data

indicate that Veph1 causes a significant increase in CTGF gene expression (Shathasivam et al.,

in preparation), which has not yet been validated, but suggests that the effect of Veph1 on

decreasing MVD is not likely due to regulation of the Hippo signalling pathway.

Decreased MVD has been associated with improved ovarian cancer survival in several

studies as it correlates with reduced intraperitoneal tumour burden. Development of metastases is

promoted by tumour vasculature providing nutrient delivery and waste removal that facilitates

rapid proliferation of tumour cells that disseminate throughout the abdomen [110-112]. In

contrast, examination of 64 primary ovarian tumours by Gadducci et al. [113] revealed that low

ovarian tumour MVD is correlated with poor chemoresponse and reduced overall survival. This

suggests that Veph1 expression may be a predictive marker of chemosensitivity.

Chemoresistance of tumours with reduced vasculature may be due to an impaired delivery of

chemotherapeutics, which is common when tumour cells are distanced from blood vessels [113].

These contrasting reports regarding the impact of MVD on survival are continued when

examining the effect of tumour cell death.

As a result of poor tumour perfusion, low MVD can result in tumour cell death often by

apoptosis or necrosis [27]. Necrosis elicits an immune response due to the rupture of necrotic cell

membranes resulting in release of cytoplasmic debris and pro-inflammatory cytokines. In

contrast, apoptosis is infrequently accompanied by an inflammatory response as apoptotic bodies

are phagocytosed and cell membranes are not ruptured [35, 36]. TUNEL assays detect DNA

fragmentation in apoptotic and necrotic cells and thus additional markers must be used to

distinguish between these two forms of cell death. Caspase-3 cleavage is an early event in

apoptosis; however, it is not likely to occur during necrosis [36, 114]. Morphology (as examined

gynecologic oncology pathologist Dr. Blaise Clarke) and double TUNEL/cleaved caspase-3

staining data indicate that observed tumour cell death is likely due to necrosis in the Veph1

expressing tumours.

Page 102: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

89

This finding has clinical implications as necrotic cell death causes a pro-inflammatory

tumour microenvironment that can be both tumour suppressing, by initiating an immune

response that causes the destruction of tumour cells, or tumour promoting, by increasing ROS

and cytokine release from infiltrating immune cells that induce DNA damage and replicative

stress of tumour cells [37]. Studies indicate that low tumour necrosis is predictive of recurrent

disease and poor survival in patients with ovarian cancer [38]; however, it is also associated with

improved prognosis of colorectal patients [39]. These contrasting reports suggest that Veph1-

dependent impairment of angiogenesis and resulting necrosis likely has clinical relevance;

however, it is not clear how this would affect overall survival.

Although we initially predicted that increased Veph1 expression would potentiate ovarian

tumour progression due to data suggesting a possible driver function of this protein as VEPH1

gene copy number is increased in a subset of ovarian cancer samples [4, 5], we discovered that

Veph1 improved survival in a SKOV3 xenograft model. Our finding is consistent with data from

The Cancer Genome Atlas indicating that VEPH1 gene variations are significantly associated

with improved survival. Genomic analysis of 311 ovarian serous cystadenomas reveals that

17.7% of tumours had a VEPH1 gene variation (54 of 55 variations were gene amplifications;

Figure 25) (http://www.cbioportal.org). These data are concordant with my results. This

suggests that an increase in VEPH1 gene copy number may have a passenger effect or be a

survival advantage. The effect of Veph1 on survival can be further investigated performing

ovarian tumour microarrays upon the availability of a suitable Veph1 antibody and using

additional mouse models that do not have the limitations in the approach used in this study.

Advantages of the subcutaneous xenograft tumour mouse model used in this study

include the ability to accurately measure tumour volume, as the tumour site is accessible, and to

evaluate the effect of Veph1 on primary tumour formation and progression. However, there are

limitations to this approach. As a human ovarian cell line was used, immunocompromised Bagg

Albino/c (BALB/c) nude mice were required to prevent a host immune response and to allow

human cell survival. This strain of mice has a Forkhead Box N1 (FoxN1) mutation causing

thymus epithelium deficiency that results in absent T cell development and impaired B cell

maturation (as described by the supplier, Charles River Laboratories). This influences the tumour

Page 103: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

90

Figure 26: Survival of serous ovarian cancer patients with and without VEPH1 alterations.

Kaplan-Meier survival curve indicating percentage of surviving serous ovarian cancer patients

with (n=55) and without (n=257) VEPH1 variations as a function of time. These variations were

predominantly gene amplifications. Data were analyzed using a Logrank Test (p<0.05).

Page 104: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

91

microenvironment by preventing the lymphocyte-mediated host tumour response that can

prevent tumour cell survival or promote inflammatory-mediated tumour progression [80]. By

using immunocompromised mice, I may not be capturing the ultimate impact of extensive

tumour necrosis in SKOV3-Ve tumours. Furthermore, the subcutaneous model used in this study

requires the placement of cells in the dorsal region of the mouse. This site of tumour cell

injection allows for accessible tumour volume monitoring; however, the ovarian cells are not in

their physiological environment of the peritoneal cavity and the impact of Veph1 on metastases

and ascites development was not addressable. As this was the first study examining the effect of

Veph1 on ovarian tumourigenesis in vivo, analysis of primary tumour progression was of initial

interest; however, future experiments exploring the effect of Veph1 on other properties of

ovarian cancer progression warrant investigation.

An additional limitation to our model is the use of only one ovarian cancer cell line that

may not be representative of a tumour histotype. Genomic profiling reveals that, although

SKOV3 is the most commonly used cell line in ovarian cancer research, these cells do not

represent any specific tumour histotype [115]. Future studies using other cell lines that more

effectively represent tumour histotypes will be beneficial in determining if Veph1 has a

histotype-specific role.

4.4 Future Directions

As my study indicates a significant impact of Veph1 on impairing angiogenesis, it is

warranted to investigate the effect of this protein on additional properties of ovarian cancer

progression. Low MVD has been correlated with both increased and decreased survival, which is

often determined by tumour chemoresponse and the extent of metastases and ascites formation

[110-113]. Veph1 has been identified as a novel inhibitor of the canonical TGF-β signalling

(Shathasivam et al., in preparation) pathway that has been implicated in promoting angiogenesis,

EMT required for metastases development, and ascites accumulation by causing abnormalization

of the diaphragm lymphatic vessel network [40]. This suggests that Veph1 may affect ascites and

metastases progression, in addition to observed effects on angiogenesis. It is thus possible that

Veph1 has a greater impact on ovarian tumour progression and survival than what has been

identified in my study; additional mouse models are required.

Page 105: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

92

Human xenograft murine models can be used to examine metastases and ascites

development when cells are injected directly into the intraperitoneal cavity or into the intrabursal

area beneath the bursal membrane on the ovarian surface. In these approaches,

immunocompromised mice would be required which would not allow for the influence of

immune responses to be included [116]. An additional mouse model of ovarian cancer involves

the use of murine ovarian cancer cells injected into non-immunodeficient mice which allows for

the host immune response to remain [117]. Tumour cell migration and metastases can be

effectively monitored by using in vivo imaging of GFP-labeled tumour cells and ascites

accumulation can be monitored by measuring abdominal girth [116]. Upon establishing the

effect of Veph1 on additional processes in ovarian tumour progress, in vitro studies will be of

importance to determine the mechanism by which Veph1 exerts its effects.

The potential impact of Veph1 on human disease had not been investigated prior to this

study. My findings indicate that Veph1 has important functions in normal physiological

conditions and in pathologies, such as cancer. I have expanded on what is known about tissue

expression of Veph1 and I have identified an interesting differential pattern of transcript isoform

expression during murine embryonic development. Additional studies are required to verify

murine Veph1A expression and localization identified in my study. I have also identified Veph1

as a novel disruptor of angiogenesis and a promoter of ‘survival’ (as defined by tumours

reaching end-point volume) in mice with ovarian xenograft tumours. Future studies to establish

the effect of Veph1 on additional properties affecting ovarian tumour progression, such as

chemosensitivity and formation of metastases and ascites, will be of great importance. Although

little is known of the specific mechanisms by which Veph1 exerts its effects, studies indicate it

may modulate multiple signalling pathways. The role of Veph1 in development, physiological

processes, and cancer progression is thus likely to be quite complex. Further work investigating

this protein is warranted and should lead to important further discoveries.

Page 106: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

93

Chapter 5

REFERENCES

1. Society, C.C. Canadian Cancer Statistics 2012. 2012; Available from:

https://www.cancer.ca/Canada-

wide/About%20cancer/Cancer%20statistics.aspx?sc_lang=EN.

2. Mantia-Smaldone, G.M., R.P. Edwards, and A.M. Vlad, Targeted treatment of recurrent

platinum-resistant ovarian cancer: current and emerging therapies. Cancer management

and research, 2011. 3: p. 25-38.

3. Herzog, T.J. and B. Pothuri, Ovarian cancer: a focus on management of recurrent

disease. Nature clinical practice. Oncology, 2006. 3(11): p. 604-11.

4. Ramakrishna, M., L.H. Williams, S.E. Boyle, J.L. Bearfoot, A. Sridhar, T.P. Speed, K.L.

Gorringe, and I.G. Campbell, Identification of candidate growth promoting genes in ovarian

cancer through integrated copy number and expression analysis. PloS one, 2010. 5(4): p.

e9983.

5. Tan, D.S., M.B. Lambros, S. Rayter, R. Natrajan, R. Vatcheva, Q. Gao, C. Marchio, F.C.

Geyer, K. Savage, S. Parry, K. Fenwick, N. Tamber, A. Mackay, T. Dexter, C. Jameson,

W.G. McCluggage, A. Williams, A. Graham, D. Faratian, M. El-Bahrawy, A.J. Paige, H.

Gabra, M.E. Gore, M. Zvelebil, C.J. Lord, S.B. Kaye, A. Ashworth, and J.S. Reis-Filho,

PPM1D is a potential therapeutic target in ovarian clear cell carcinomas. Clinical cancer

research, 2009. 15(7): p. 2269-80.

6. Sodek, K.L., M.J. Ringuette, and T.J. Brown, MT1-MMP is the critical determinant of

matrix degradation and invasion by ovarian cancer cells. British journal of cancer, 2007.

97(3): p. 358-67.

7. Sodek, K.L., M.J. Ringuette, and T.J. Brown, Compact spheroid formation by ovarian

cancer cells is associated with contractile behavior and an invasive phenotype. International

journal of cancer, 2009. 124(9): p. 2060-70.

8. Teleman, A.A., Y.W. Chen, and S.M. Cohen, Drosophila Melted modulates FOXO and

TOR activity. Developmental cell, 2005. 9(2): p. 271-81.

9. Jukam, D. and C. Desplan, Binary regulation of Hippo pathway by Merlin/NF2, Kibra,

Lgl, and Melted specifies and maintains postmitotic neuronal fate. Developmental cell,

2011. 21(5): p. 874-87.

10. Mikeladze-Dvali, T., M.F. Wernet, D. Pistillo, E.O. Mazzoni, A.A. Teleman, Y.W. Chen,

S. Cohen, and C. Desplan, The growth regulators warts/lats and melted interact in a bistable

loop to specify opposite fates in Drosophila R8 photoreceptors. Cell, 2005. 122(5): p. 775-

87.

11. Hsiao, H.Y., D. Jukam, R. Johnston, and C. Desplan, The neuronal transcription factor

erect wing regulates specification and maintenance of Drosophila R8 photoreceptor

subtypes. Developmental biology, 2013. 381(2): p. 482-90.

12. Muto, E., Y. Tabata, T. Taneda, Y. Aoki, A. Muto, K. Arai, and S. Watanabe,

Identification and characterization of Veph, a novel gene encoding a PH domain-containing

protein expressed in the developing central nervous system of vertebrates. Biochimie, 2004.

86(8): p. 523-31.

13. Jemal, A., R. Siegel, E. Ward, Y. Hao, J. Xu, T. Murray, and M.J. Thun, Cancer

statistics, 2008. CA: a cancer journal for clinicians, 2008. 58(2): p. 71-96.

Page 107: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

94

14. Sankaranarayanan, R. and J. Ferlay, Worldwide burden of gynaecological cancer: the

size of the problem. Best practice & research. Clinical obstetrics & gynaecology, 2006.

20(2): p. 207-25.

15. Kim, J., D.M. Coffey, C.J. Creighton, Z. Yu, S.M. Hawkins, and M.M. Matzuk, High-

grade serous ovarian cancer arises from fallopian tube in a mouse model. Proceedings of

the National Academy of Sciences of the United States of America, 2012. 109(10): p. 3921-

6.

16. Karst, A.M. and R. Drapkin, Ovarian cancer pathogenesis: a model in evolution. Journal

of oncology, 2010. 2010: p. 932371.

17. Levanon, K., C. Crum, and R. Drapkin, New insights into the pathogenesis of serous

ovarian cancer and its clinical impact. Journal of clinical oncology, 2008. 26(32): p. 5284-

93.

18. Goff, B.A., Ovarian cancer: screening and early detection. Obstetrics and gynecology

clinics of North America, 2012. 39(2): p. 183-94.

19. Bast, R.C., Jr., B. Hennessy, and G.B. Mills, The biology of ovarian cancer: new

opportunities for translation. Nature reviews. Cancer, 2009. 9(6): p. 415-28.

20. Chen, V.W., B. Ruiz, J.L. Killeen, T.R. Cote, X.C. Wu, and C.N. Correa, Pathology and

classification of ovarian tumors. Cancer, 2003. 97(10 Suppl): p. 2631-42.

21. Kurman, R.J. and M. Shih Ie, Molecular pathogenesis and extraovarian origin of

epithelial ovarian cancer--shifting the paradigm. Human pathology, 2011. 42(7): p. 918-31.

22. Seidman, J.D., I. Horkayne-Szakaly, M. Haiba, C.R. Boice, R.J. Kurman, and B.M.

Ronnett, The histologic type and stage distribution of ovarian carcinomas of surface

epithelial origin. International journal of gynecological pathology, 2004. 23(1): p. 41-4.

23. Lalwani, N., S.R. Prasad, R. Vikram, A.K. Shanbhogue, P.C. Huettner, and N. Fasih,

Histologic, molecular, and cytogenetic features of ovarian cancers: implications for

diagnosis and treatment. Radiographics : a review publication of the Radiological Society of

North America, Inc, 2011. 31(3): p. 625-46.

24. Integrated genomic analyses of ovarian carcinoma. Nature, 2011. 474(7353): p. 609-15.

25. Tothill, R.W., A.V. Tinker, J. George, R. Brown, S.B. Fox, S. Lade, D.S. Johnson, M.K.

Trivett, D. Etemadmoghadam, B. Locandro, N. Traficante, S. Fereday, J.A. Hung, Y.E.

Chiew, I. Haviv, D. Gertig, A. DeFazio, and D.D. Bowtell, Novel molecular subtypes of

serous and endometrioid ovarian cancer linked to clinical outcome. Clinical cancer research,

2008. 14(16): p. 5198-208.

26. del Carmen, M.G., M. Birrer, and J.O. Schorge, Clear cell carcinoma of the ovary: a

review of the literature. Gynecologic oncology, 2012. 126(3): p. 481-90.

27. Bamberger, E.S. and C.W. Perrett, Angiogenesis in epithelian ovarian cancer. Molecular

pathology : MP, 2002. 55(6): p. 348-59.

28. Lengyel, E., Ovarian cancer development and metastasis. The American journal of

pathology, 2010. 177(3): p. 1053-64.

29. Tan, D.S., R. Agarwal, and S.B. Kaye, Mechanisms of transcoelomic metastasis in

ovarian cancer. The lancet oncology, 2006. 7(11): p. 925-34.

30. Frankel, A., K. Rosen, J. Filmus, and R.S. Kerbel, Induction of anoikis and suppression

of human ovarian tumor growth in vivo by down-regulation of Bcl-X(L). Cancer research,

2001. 61(12): p. 4837-41.

31. Tang, M.K., H.Y. Zhou, J.W. Yam, and A.S. Wong, c-Met overexpression contributes to

the acquired apoptotic resistance of nonadherent ovarian cancer cells through a cross talk

Page 108: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

95

mediated by phosphatidylinositol 3-kinase and extracellular signal-regulated kinase 1/2.

Neoplasia, 2010. 12(2): p. 128-38.

32. Sodek, K.L., K.J. Murphy, T.J. Brown, and M.J. Ringuette, Cell-cell and cell-matrix

dynamics in intraperitoneal cancer metastasis. Cancer metastasis reviews, 2012. 31(1-2): p.

397-414.

33. Lebrin, F., M. Deckers, P. Bertolino, and P. Ten Dijke, TGF-beta receptor function in the

endothelium. Cardiovascular research, 2005. 65(3): p. 599-608.

34. Potenta, S., E. Zeisberg, and R. Kalluri, The role of endothelial-to-mesenchymal

transition in cancer progression. British journal of cancer, 2008. 99(9): p. 1375-9.

35. Elmore, S., Apoptosis: a review of programmed cell death. Toxicologic pathology, 2007.

35(4): p. 495-516.

36. Zong, W.X. and C.B. Thompson, Necrotic death as a cell fate. Genes & development,

2006. 20(1): p. 1-15.

37. Vakkila, J. and M.T. Lotze, Inflammation and necrosis promote tumour growth. Nature

reviews. Immunology, 2004. 4(8): p. 641-8.

38. Le, T., P. Shahriari, L. Hopkins, W. Faught, and M. Fung Kee Fung, Prognostic

significance of tumor necrosis in ovarian cancer patients treated with neoadjuvant

chemotherapy and interval surgical debulking. International journal of gynecological cancer,

2006. 16(3): p. 986-90.

39. Richards, C.H., C.S. Roxburgh, J.H. Anderson, R.F. McKee, A.K. Foulis, P.G. Horgan,

and D.C. McMillan, Prognostic value of tumour necrosis and host inflammatory responses

in colorectal cancer. The British journal of surgery, 2012. 99(2): p. 287-94.

40. Liao, S., J. Liu, P. Lin, T. Shi, R.K. Jain, and L. Xu, TGF-beta blockade controls ascites

by preventing abnormalization of lymphatic vessels in orthotopic human ovarian carcinoma

models. Clinical cancer research, 2011. 17(6): p. 1415-24.

41. Kipps, E., D.S. Tan, and S.B. Kaye, Meeting the challenge of ascites in ovarian cancer:

new avenues for therapy and research. Nature reviews. Cancer, 2013. 13(4): p. 273-82.

42. Woopen, H. and J. Sehouli, Current and future options in the treatment of malignant

ascites in ovarian cancer. Anticancer research, 2009. 29(8): p. 3353-9.

43. Maccio, A. and C. Madeddu, Inflammation and ovarian cancer. Cytokine, 2012. 58(2): p.

133-47.

44. Richards, J.S., D.L. Russell, R.L. Robker, M. Dajee, and T.N. Alliston, Molecular

mechanisms of ovulation and luteinization. Molecular and cellular endocrinology, 1998.

145(1-2): p. 47-54.

45. Shkolnik, K., A. Tadmor, S. Ben-Dor, N. Nevo, D. Galiani, and N. Dekel, Reactive

oxygen species are indispensable in ovulation. Proceedings of the National Academy of

Sciences of the United States of America, 2011. 108(4): p. 1462-7.

46. Szabova, L., C. Yin, S. Bupp, T.M. Guerin, J.J. Schlomer, D.B. Householder, M.L.

Baran, M. Yi, Y. Song, W. Sun, J.E. McDunn, P.L. Martin, T. Van Dyke, and S.

Difilippantonio, Perturbation of Rb, p53, and Brca1 or Brca2 cooperate in inducing

metastatic serous epithelial ovarian cancer. Cancer research, 2012. 72(16): p. 4141-53.

47. Flesken-Nikitin, A., C.I. Hwang, C.Y. Cheng, T.V. Michurina, G. Enikolopov, and A.Y.

Nikitin, Ovarian surface epithelium at the junction area contains a cancer-prone stem cell

niche. Nature, 2013. 495(7440): p. 241-5.

Page 109: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

96

48. Maines-Bandiera, S.L. and N. Auersperg, Increased E-cadherin expression in ovarian

surface epithelium: an early step in metaplasia and dysplasia? International journal of

gynecological pathology, 1997. 16(3): p. 250-5.

49. King, S.M., T.S. Hilliard, L.Y. Wu, R.C. Jaffe, A.T. Fazleabas, and J.E. Burdette, The

impact of ovulation on fallopian tube epithelial cells: evaluating three hypotheses

connecting ovulation and serous ovarian cancer. Endocrine-related cancer, 2011. 18(5): p.

627-42.

50. Karst, A.M., K. Levanon, and R. Drapkin, Modeling high-grade serous ovarian

carcinogenesis from the fallopian tube. Proceedings of the National Academy of Sciences of

the United States of America, 2011. 108(18): p. 7547-52.

51. Marquez, R.T., K.A. Baggerly, A.P. Patterson, J. Liu, R. Broaddus, M. Frumovitz, E.N.

Atkinson, D.I. Smith, L. Hartmann, D. Fishman, A. Berchuck, R. Whitaker, D.M.

Gershenson, G.B. Mills, R.C. Bast, Jr., and K.H. Lu, Patterns of gene expression in different

histotypes of epithelial ovarian cancer correlate with those in normal fallopian tube,

endometrium, and colon. Clinical cancer research, 2005. 11(17): p. 6116-26.

52. Tone, A.A., H. Begley, M. Sharma, J. Murphy, B. Rosen, T.J. Brown, and P.A. Shaw,

Gene expression profiles of luteal phase fallopian tube epithelium from BRCA mutation

carriers resemble high-grade serous carcinoma. Clinical cancer research, 2008. 14(13): p.

4067-78.

53. Tone, A.A., C. Virtanen, P. Shaw, and T.J. Brown, Prolonged postovulatory

proinflammatory signaling in the fallopian tube epithelium may be mediated through a

BRCA1/DAB2 axis. Clinical cancer research, 2012. 18(16): p. 4334-44.

54. Medeiros, F., M.G. Muto, Y. Lee, J.A. Elvin, M.J. Callahan, C. Feltmate, J.E. Garber,

D.W. Cramer, and C.P. Crum, The tubal fimbria is a preferred site for early

adenocarcinoma in women with familial ovarian cancer syndrome. The American journal of

surgical pathology, 2006. 30(2): p. 230-6.

55. Callahan, M.J., C.P. Crum, F. Medeiros, D.W. Kindelberger, J.A. Elvin, J.E. Garber,

C.M. Feltmate, R.S. Berkowitz, and M.G. Muto, Primary fallopian tube malignancies in

BRCA-positive women undergoing surgery for ovarian cancer risk reduction. Journal of

clinical oncology, 2007. 25(25): p. 3985-90.

56. Kobayashi, H., H. Kajiwara, S. Kanayama, Y. Yamada, N. Furukawa, T. Noguchi, S.

Haruta, S. Yoshida, M. Sakata, T. Sado, and H. Oi, Molecular pathogenesis of

endometriosis-associated clear cell carcinoma of the ovary (review). Oncology reports,

2009. 22(2): p. 233-40.

57. Robboy, S.J. and S.M. Bean, Pathogenesis of endometriosis. Reproductive biomedicine

online, 2010. 21(1): p. 4-5.

58. Stany, M.P., V. Vathipadiekal, L. Ozbun, R.L. Stone, S.C. Mok, H. Xue, T. Kagami, Y.

Wang, J.N. McAlpine, D. Bowtell, P.W. Gout, D.M. Miller, C.B. Gilks, D.G. Huntsman,

S.L. Ellard, Y.Z. Wang, P. Vivas-Mejia, G. Lopez-Berestein, A.K. Sood, and M.J. Birrer,

Identification of novel therapeutic targets in microdissected clear cell ovarian cancers. PloS

one, 2011. 6(7): p. e21121.

59. Itamochi, H., J. Kigawa, T. Sugiyama, Y. Kikuchi, M. Suzuki, and N. Terakawa, Low

proliferation activity may be associated with chemoresistance in clear cell carcinoma of the

ovary. Obstetrics and gynecology, 2002. 100(2): p. 281-7.

Page 110: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

97

60. Prokopenko, S.N., Y. He, Y. Lu, and H.J. Bellen, Mutations affecting the development of

the peripheral nervous system in Drosophila: a molecular screen for novel proteins.

Genetics, 2000. 156(4): p. 1691-715.

61. Salzberg, A., S.N. Prokopenko, Y. He, P. Tsai, M. Pal, P. Maroy, D.M. Glover, P. Deak,

and H.J. Bellen, P-element insertion alleles of essential genes on the third chromosome of

Drosophila melanogaster: mutations affecting embryonic PNS development. Genetics, 1997.

147(4): p. 1723-41.

62. Rebecchi, M.J. and S. Scarlata, Pleckstrin homology domains: a common fold with

diverse functions. Annual review of biophysics and biomolecular structure, 1998. 27: p. 503-

28.

63. Scheffzek, K. and S. Welti, Pleckstrin homology (PH) like domains - versatile modules in

protein-protein interaction platforms. FEBS letters, 2012. 586(17): p. 2662-73.

64. Barrios-Rodiles, M., K.R. Brown, B. Ozdamar, R. Bose, Z. Liu, R.S. Donovan, F. Shinjo,

Y. Liu, J. Dembowy, I.W. Taylor, V. Luga, N. Przulj, M. Robinson, H. Suzuki, Y.

Hayashizaki, I. Jurisica, and J.L. Wrana, High-throughput mapping of a dynamic signaling

network in mammalian cells. Science, 2005. 307(5715): p. 1621-5.

65. Tran, H., A. Brunet, E.C. Griffith, and M.E. Greenberg, The many forks in FOXO's road.

Science's STKE : signal transduction knowledge environment, 2003. 2003(172): p. RE5.

66. Tran, H., A. Brunet, J.M. Grenier, S.R. Datta, A.J. Fornace, Jr., P.S. DiStefano, L.W.

Chiang, and M.E. Greenberg, DNA repair pathway stimulated by the forkhead transcription

factor FOXO3a through the Gadd45 protein. Science, 2002. 296(5567): p. 530-4.

67. Sunayama, J., F. Tsuruta, N. Masuyama, and Y. Gotoh, JNK antagonizes Akt-mediated

survival signals by phosphorylating 14-3-3. The Journal of cell biology, 2005. 170(2): p.

295-304.

68. Wang, Y., L. Dong, H. Cui, D.H. Shen, X.H. Chang, T.Y. Fu, X. Ye, and Y.Y. Yao, Up-

regulation of mitochondrial antioxidation signals in ovarian cancer cells with aggressive

biologic behavior. Journal of Zhejiang University. Science. B, 2011. 12(5): p. 346-56.

69. Connor, K.M., N. Hempel, K.K. Nelson, G. Dabiri, A. Gamarra, J. Belarmino, L. Van De

Water, B.M. Mian, and J.A. Melendez, Manganese superoxide dismutase enhances the

invasive and migratory activity of tumor cells. Cancer research, 2007. 67(21): p. 10260-7.

70. Storz, P., H. Doppler, J.A. Copland, K.J. Simpson, and A. Toker, FOXO3a promotes

tumor cell invasion through the induction of matrix metalloproteinases. Molecular and

cellular biology, 2009. 29(18): p. 4906-17.

71. Connor, K.M., S. Subbaram, K.J. Regan, K.K. Nelson, J.E. Mazurkiewicz, P.J.

Bartholomew, A.E. Aplin, Y.T. Tai, J. Aguirre-Ghiso, S.C. Flores, and J.A. Melendez,

Mitochondrial H2O2 regulates the angiogenic phenotype via PTEN oxidation. The Journal

of biological chemistry, 2005. 280(17): p. 16916-24.

72. Potente, M., C. Urbich, K. Sasaki, W.K. Hofmann, C. Heeschen, A. Aicher, R. Kollipara,

R.A. DePinho, A.M. Zeiher, and S. Dimmeler, Involvement of Foxo transcription factors in

angiogenesis and postnatal neovascularization. The Journal of clinical investigation, 2005.

115(9): p. 2382-92.

73. Reagan-Shaw, S. and N. Ahmad, The role of Forkhead-box Class O (FoxO) transcription

factors in cancer: a target for the management of cancer. Toxicology and applied

pharmacology, 2007. 224(3): p. 360-8.

Page 111: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

98

74. Karger, S., C. Weidinger, K. Krause, S.Y. Sheu, T. Aigner, O. Gimm, K.W. Schmid, H.

Dralle, and D. Fuhrer, FOXO3a: a novel player in thyroid carcinogenesis? Endocrine-

related cancer, 2009. 16(1): p. 189-99.

75. Lu, M., J. Xiang, F. Xu, Y. Wang, Y. Yin, and D. Chen, The expression and significance

of pThr32-FOXO3a in human ovarian cancer. Medical oncology, 2012. 29(2): p. 1258-64.

76. Fei, M., Y. Zhao, Y. Wang, M. Lu, C. Cheng, X. Huang, D. Zhang, J. Lu, S. He, and A.

Shen, Low expression of Foxo3a is associated with poor prognosis in ovarian cancer

patients. Cancer investigation, 2009. 27(1): p. 52-9.

77. Chen, J., A.R. Gomes, L.J. Monteiro, S.Y. Wong, L.H. Wu, T.T. Ng, C.T. Karadedou, J.

Millour, Y.C. Ip, Y.N. Cheung, A. Sunters, K.Y. Chan, E.W. Lam, and U.S. Khoo,

Constitutively nuclear FOXO3a localization predicts poor survival and promotes Akt

phosphorylation in breast cancer. PloS one, 2010. 5(8): p. e12293.

78. Tenbaum, S.P., P. Ordonez-Moran, I. Puig, I. Chicote, O. Arques, S. Landolfi, Y.

Fernandez, J.R. Herance, J.D. Gispert, L. Mendizabal, S. Aguilar, S. Ramon y Cajal, S.

Schwartz, Jr., A. Vivancos, E. Espin, S. Rojas, J. Baselga, J. Tabernero, A. Munoz, and H.G.

Palmer, beta-catenin confers resistance to PI3K and AKT inhibitors and subverts FOXO3a

to promote metastasis in colon cancer. Nature medicine, 2012. 18(6): p. 892-901.

79. Huang, J., L. Zhang, J. Greshock, T.A. Colligon, Y. Wang, R. Ward, D. Katsaros, H.

Lassus, R. Butzow, A.K. Godwin, J.R. Testa, K.L. Nathanson, P.A. Gimotty, G. Coukos,

B.L. Weber, and Y. Degenhardt, Frequent genetic abnormalities of the PI3K/AKT pathway

in primary ovarian cancer predict patient outcome. Genes, chromosomes & cancer, 2011.

50(8): p. 606-18.

80. Massague, J., TGFbeta in Cancer. Cell, 2008. 134(2): p. 215-30.

81. Seoane, J., H.V. Le, L. Shen, S.A. Anderson, and J. Massague, Integration of Smad and

forkhead pathways in the control of neuroepithelial and glioblastoma cell proliferation.

Cell, 2004. 117(2): p. 211-23.

82. Davidson, B., C.G. Trope, and R. Reich, Epithelial-mesenchymal transition in ovarian

carcinoma. Frontiers in oncology, 2012. 2: p. 33.

83. Medici, D., E.D. Hay, and B.R. Olsen, Snail and Slug promote epithelial-mesenchymal

transition through beta-catenin-T-cell factor-4-dependent expression of transforming growth

factor-beta3. Molecular biology of the cell, 2008. 19(11): p. 4875-87.

84. Haslehurst, A.M., M. Koti, M. Dharsee, P. Nuin, K. Evans, J. Geraci, T. Childs, J. Chen,

J. Li, J. Weberpals, S. Davey, J. Squire, P.C. Park, and H. Feilotter, EMT transcription

factors snail and slug directly contribute to cisplatin resistance in ovarian cancer. BMC

cancer, 2012. 12: p. 91.

85. Rodriguez, G.C., C. Haisley, J. Hurteau, T.L. Moser, R. Whitaker, R.C. Bast, Jr., and

M.S. Stack, Regulation of invasion of epithelial ovarian cancer by transforming growth

factor-beta. Gynecologic oncology, 2001. 80(2): p. 245-53.

86. Medici, D., S. Potenta, and R. Kalluri, Transforming growth factor-beta2 promotes Snail-

mediated endothelial-mesenchymal transition through convergence of Smad-dependent and

Smad-independent signalling. The Biochemical journal, 2011. 437(3): p. 515-20.

87. Zeisberg, E.M., S. Potenta, L. Xie, M. Zeisberg, and R. Kalluri, Discovery of endothelial

to mesenchymal transition as a source for carcinoma-associated fibroblasts. Cancer

research, 2007. 67(21): p. 10123-8.

88. Luga, V., L. Zhang, A.M. Viloria-Petit, A.A. Ogunjimi, M.R. Inanlou, E. Chiu, M.

Buchanan, A.N. Hosein, M. Basik, and J.L. Wrana, Exosomes mediate stromal mobilization

Page 112: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

99

of autocrine Wnt-PCP signaling in breast cancer cell migration. Cell, 2012. 151(7): p. 1542-

56.

89. Pan, D., The hippo signaling pathway in development and cancer. Developmental cell,

2010. 19(4): p. 491-505.

90. Hall, C.A., R. Wang, J. Miao, E. Oliva, X. Shen, T. Wheeler, S.G. Hilsenbeck, S. Orsulic,

and S. Goode, Hippo pathway effector Yap is an ovarian cancer oncogene. Cancer research,

2010. 70(21): p. 8517-25.

91. Zhang, X., J. George, S. Deb, J.L. Degoutin, E.A. Takano, S.B. Fox, D.D. Bowtell, and

K.F. Harvey, The Hippo pathway transcriptional co-activator, YAP, is an ovarian cancer

oncogene. Oncogene, 2011. 30(25): p. 2810-22.

92. Huntoon, C.J., M.D. Nye, L. Geng, K.L. Peterson, K.S. Flatten, P. Haluska, S.H.

Kaufmann, and L.M. Karnitz, Heat shock protein 90 inhibition depletes LATS1 and LATS2,

two regulators of the mammalian hippo tumor suppressor pathway. Cancer research, 2010.

70(21): p. 8642-50.

93. Calvo, F., N. Ege, A. Grande-Garcia, S. Hooper, R.P. Jenkins, S.I. Chaudhry, K.

Harrington, P. Williamson, E. Moeendarbary, G. Charras, and E. Sahai,

Mechanotransduction and YAP-dependent matrix remodelling is required for the generation

and maintenance of cancer-associated fibroblasts. Nature cell biology, 2013. 15(6): p. 637-

46.

94. Sjoblom, T., S. Jones, L.D. Wood, D.W. Parsons, J. Lin, T.D. Barber, D. Mandelker, R.J.

Leary, J. Ptak, N. Silliman, S. Szabo, P. Buckhaults, C. Farrell, P. Meeh, S.D. Markowitz, J.

Willis, D. Dawson, J.K. Willson, A.F. Gazdar, J. Hartigan, L. Wu, C. Liu, G. Parmigiani,

B.H. Park, K.E. Bachman, N. Papadopoulos, B. Vogelstein, K.W. Kinzler, and V.E.

Velculescu, The consensus coding sequences of human breast and colorectal cancers.

Science, 2006. 314(5797): p. 268-74.

95. Hernandez, P., X. Sole, J. Valls, V. Moreno, G. Capella, A. Urruticoechea, and M.A.

Pujana, Integrative analysis of a cancer somatic mutome. Molecular cancer, 2007. 6: p. 13.

96. Ylipaa, A., K.K. Hunt, J. Yang, A.J. Lazar, K.E. Torres, D.C. Lev, M. Nykter, R.E.

Pollock, J. Trent, and W. Zhang, Integrative genomic characterization and a genomic

staging system for gastrointestinal stromal tumors. Cancer, 2011. 117(2): p. 380-9.

97. Pharoah, P.D., Y.Y. Tsai, S.J. Ramus, C.M. Phelan, E.L. Goode, K. Lawrenson, M.

Buckley, B.L. Fridley, J.P. Tyrer, H. Shen, R. Weber, R. Karevan, M.C. Larson, H. Song,

D.C. Tessier, F. Bacot, D. Vincent, J.M. Cunningham, J. Dennis, E. Dicks, K.K. Aben, H.

Anton-Culver, N. Antonenkova, S.M. Armasu, L. Baglietto, E.V. Bandera, M.W.

Beckmann, M.J. Birrer, G. Bloom, N. Bogdanova, J.D. Brenton, L.A. Brinton, A. Brooks-

Wilson, R. Brown, R. Butzow, I. Campbell, M.E. Carney, R.S. Carvalho, J. Chang-Claude,

Y.A. Chen, Z. Chen, W.H. Chow, M.S. Cicek, G. Coetzee, L.S. Cook, D.W. Cramer, C.

Cybulski, A. Dansonka-Mieszkowska, E. Despierre, J.A. Doherty, T. Dork, A. du Bois, M.

Durst, D. Eccles, R. Edwards, A.B. Ekici, P.A. Fasching, D. Fenstermacher, J. Flanagan,

Y.T. Gao, M. Garcia-Closas, A. Gentry-Maharaj, G. Giles, A. Gjyshi, M. Gore, J. Gronwald,

Q. Guo, M.K. Halle, P. Harter, A. Hein, F. Heitz, P. Hillemanns, M. Hoatlin, E. Hogdall,

C.K. Hogdall, S. Hosono, A. Jakubowska, A. Jensen, K.R. Kalli, B.Y. Karlan, L.E.

Kelemen, L.A. Kiemeney, S.K. Kjaer, G.E. Konecny, C. Krakstad, J. Kupryjanczyk, D.

Lambrechts, S. Lambrechts, N.D. Le, N. Lee, J. Lee, A. Leminen, B.K. Lim, J. Lissowska, J.

Lubinski, L. Lundvall, G. Lurie, L.F. Massuger, K. Matsuo, V. McGuire, J.R. McLaughlin,

U. Menon, F. Modugno, K.B. Moysich, T. Nakanishi, S.A. Narod, R.B. Ness, H.

Page 113: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

100

Nevanlinna, S. Nickels, H. Noushmehr, K. Odunsi, S. Olson, I. Orlow, J. Paul, T. Pejovic,

L.M. Pelttari, J. Permuth-Wey, M.C. Pike, E.M. Poole, X. Qu, H.A. Risch, L. Rodriguez-

Rodriguez, M.A. Rossing, A. Rudolph, I. Runnebaum, I.K. Rzepecka, H.B. Salvesen, I.

Schwaab, G. Severi, V. Shridhar, X.O. Shu, W. Sieh, M.C. Southey, P. Spellman, K. Tajima,

S.H. Teo, K.L. Terry, P.J. Thompson, A. Timorek, S.S. Tworoger, A.M. van Altena, D. van

den Berg, I. Vergote, R.A. Vierkant, A.F. Vitonis, S. Wang-Gohrke, N. Wentzensen, A.S.

Whittemore, E. Wik, B. Winterhoff, Y.L. Woo, A.H. Wu, H.P. Yang, W. Zheng, A. Ziogas,

F. Zulkifli, M.T. Goodman, P. Hall, D.F. Easton, C.L. Pearce, A. Berchuck, G. Chenevix-

Trench, E. Iversen, A.N. Monteiro, S.A. Gayther, J.M. Schildkraut and T.A. Sellers, GWAS

meta-analysis and replication identifies three new susceptibility loci for ovarian cancer.

Nature genetics, 2013. 45(4): p. 362-70.

98. Kollara, A. and T.J. Brown, Variable expression of nuclear receptor coactivator 4

(NcoA4) during mouse embryonic development. The journal of histochemistry and

cytochemistry, 2010. 58(7): p. 595-609.

99. Tomayko, M.M. and C.P. Reynolds, Determination of subcutaneous tumor size in

athymic (nude) mice. Cancer chemotherapy and pharmacology, 1989. 24(3): p. 148-54.

100. Kollara, A. and T.J. Brown, Functional interaction of nuclear receptor coactivator 4 with

aryl hydrocarbon receptor. Biochemical and biophysical research communications, 2006.

346(2): p. 526-34.

101. Tuominen, V.J., S. Ruotoistenmaki, A. Viitanen, M. Jumppanen, and J. Isola,

ImmunoRatio: a publicly available web application for quantitative image analysis of

estrogen receptor (ER), progesterone receptor (PR), and Ki-67. Breast cancer research :

BCR, 2010. 12(4): p. R56.

102. Vermeulen, P.B., G. Gasparini, S.B. Fox, C. Colpaert, L.P. Marson, M. Gion, J.A. Belien,

R.M. de Waal, E. Van Marck, E. Magnani, N. Weidner, A.L. Harris, and L.Y. Dirix, Second

international consensus on the methodology and criteria of evaluation of angiogenesis

quantification in solid human tumours. European journal of cancer, 2002. 38(12): p. 1564-

79.

103. Goertz, M.J., Z. Wu, T.D. Gallardo, F.K. Hamra, and D.H. Castrillon, Foxo1 is required

in mouse spermatogonial stem cells for their maintenance and the initiation of

spermatogenesis. The Journal of clinical investigation, 2011. 121(9): p. 3456-66.

104. Sui, X.X., L.L. Luo, J.J. Xu, and Y.C. Fu, Evidence that FOXO3a is involved in oocyte

apoptosis in the neonatal rat ovary. Biochemistry and cell biology, 2010. 88(4): p. 621-8.

105. Levanon, K., V. Ng, H.Y. Piao, Y. Zhang, M.C. Chang, M.H. Roh, D.W. Kindelberger,

M.S. Hirsch, C.P. Crum, J.A. Marto, and R. Drapkin, Primary ex vivo cultures of human

fallopian tube epithelium as a model for serous ovarian carcinogenesis. Oncogene, 2010.

29(8): p. 1103-13.

106. Sundfeldt, K., Y. Piontkewitz, K. Ivarsson, O. Nilsson, P. Hellberg, M. Brannstrom, P.O.

Janson, S. Enerback, and L. Hedin, E-cadherin expression in human epithelial ovarian

cancer and normal ovary. International journal of cancer, 1997. 74(3): p. 275-80.

107. Veatch, A.L., L.F. Carson, and S. Ramakrishnan, Differential expression of the cell-cell

adhesion molecule E-cadherin in ascites and solid human ovarian tumor cells. International

journal of cancer, 1994. 58(3): p. 393-9.

108. Dong, L.L., L. Liu, C.H. Ma, J.S. Li, C. Du, S. Xu, L.H. Han, L. Li, and X.W. Wang, E-

cadherin promotes proliferation of human ovarian cancer cells in vitro via activating

MEK/ERK pathway. Acta pharmacologica Sinica, 2012. 33(6): p. 817-22.

Page 114: THE IMPACT OF VEPH1 ON OVARIAN CANCER TUMOUR …...Thomasina Spybey, Master of Science Graduate Department of Physiology, University of Toronto Toronto, Ontario, Canada 2013 ABSTRACT

101

109. Duncan, T.J., A. Al-Attar, P. Rolland, I.V. Scott, S. Deen, D.T. Liu, I. Spendlove, and

L.G. Durrant, Vascular endothelial growth factor expression in ovarian cancer: a model for

targeted use of novel therapies? Clinical cancer research, 2008. 14(10): p. 3030-5.

110. Stone, P.J., M.J. Goodheart, S.L. Rose, B.J. Smith, B.R. DeYoung, and R.E. Buller, The

influence of microvessel density on ovarian carcinogenesis. Gynecologic oncology, 2003.

90(3): p. 566-71.

111. Alvarez, A.A., H.R. Krigman, R.S. Whitaker, R.K. Dodge, and G.C. Rodriguez, The

prognostic significance of angiogenesis in epithelial ovarian carcinoma. Clinical cancer

research, 1999. 5(3): p. 587-91.

112. Nadkarni, N.J., K.D. Geest, T. Neff, B.D. Young, D.P. Bender, A. Ahmed, B.J. Smith, A.

Button, and M.J. Goodheart, Microvessel density and p53 mutations in advanced-stage

epithelial ovarian cancer. Cancer letters, 2013. 331(1): p. 99-104.

113. Gadducci, A., P. Viacava, S. Cosio, G. Fanelli, A. Fanucchi, D. Cecchetti, R. Cristofani,

and A.R. Genazzani, Intratumoral microvessel density, response to chemotherapy and

clinical outcome of patients with advanced ovarian carcinoma. Anticancer research, 2003.

23(1B): p. 549-56.

114. Hetz, C.A., M. Hunn, P. Rojas, V. Torres, L. Leyton, and A.F. Quest, Caspase-dependent

initiation of apoptosis and necrosis by the Fas receptor in lymphoid cells: onset of necrosis

is associated with delayed ceramide increase. Journal of cell science, 2002. 115(Pt 23): p.

4671-83.

115. Domcke, S., R. Sinha, D.A. Levine, C. Sander, and N. Schultz, Evaluating cell lines as

tumour models by comparison of genomic profiles. Nature communications, 2013. 4: p.

2126.

116. Sale, S. and S. Orsulic, Models of ovarian cancer metastasis: Murine models. Drug

discovery today. Disease models, 2006. 3(2): p. 149-154.

117. Greenaway, J., R. Moorehead, P. Shaw, and J. Petrik, Epithelial-stromal interaction

increases cell proliferation, survival and tumorigenicity in a mouse model of human

epithelial ovarian cancer. Gynecologic oncology, 2008. 108(2): p. 385-94.