discovery of novel ovarian cancer biomarkers via proteomics and mass

205
DISCOVERY OF NOVEL OVARIAN CANCER BIOMARKERS VIA PROTEOMICS AND MASS SPECTROMETRY By Chinthaka Geeth Gunawardana A thesis submitted in conformity with the requirements for the Degree of Doctor of Philosophy Graduate Department of Laboratory Medicine and Pathobiology University of Toronto © Copyright by Chinthaka Geeth Gunawardana 2010

Upload: others

Post on 09-Feb-2022

3 views

Category:

Documents


0 download

TRANSCRIPT

DISCOVERY OF NOVEL OVARIAN CANCER BIOMARKERS VIA PROTEOMICS

AND MASS SPECTROMETRY

By

Chinthaka Geeth Gunawardana

A thesis submitted in conformity with the requirements for

the Degree of Doctor of Philosophy

Graduate Department of Laboratory Medicine and Pathobiology

University of Toronto

© Copyright by Chinthaka Geeth Gunawardana 2010

ii

DISCOVERY OF NOVEL OVARIAN CANCER BIOMARKERS VIA PROTEOMICS

AND MASS SPECTROMETRY

Chinthaka Geeth Gunawardana

Doctor of Philosophy 2010

Department of Laboratory Medicine and Pathobiology

University of Toronto

ABSTRACT

Proteins secreted or shed by tumors can be found in serum. Detecting these proteins by

mass spectrometry (MS) is difficult, due to the wide dynamic range of protein concentrations in

serum. To circumvent this issue, we mined the conditioned media of epithelial ovarian cancer

(EOC) cell lines which is a less complex fluid to work with. We hypothesize that some of the

proteins shed or secreted by EOC cell lines are similar to those secreted or shed by EOC tumors

and that some of these proteins can be used as biomarkers. We mined the conditioned medium

of four ovarian cancer cell lines (HTB75, TOV-112D, TOV-21G and RMUG-S) by two-

dimensional liquid chromatography-mass spectrometry. Our study identified 1208, 1252, 885,

and 463 proteins from the HTB-75, TOV-112D, TOV-21G, and RMUG-S cell lines

respectively. In all, we identified 2039 proteins from which we focused on 420 extracellular and

plasma membrane proteins. High abundance proteins such as albumin and immunoglobulins,

which are problematic for serum proteomics, did not interfere with our study. Several known

markers of EOC including CA-125, HE4, Mesothelin, and KLK6, were identified in this study.

The list of 420 extracellular and membrane proteins was cross-referenced with the proteome of

ascites fluid to generate a final list of 51 potential candidates. According to Ingenuity Pathway

Analysis, two of the top 10 diseases associated with our list of 51 proteins were cancer and

iii

reproductive diseases. Of the 51 candidates, 10 proteins were selected for verification in sera

from ovarian cancer patients and healthy individuals. Clusterin showed a significant difference

between cancer patients and normal, with sera from cancer patients showing higher levels.

Another protein, NPC2, did not show a difference in sera between cancer and normals. Protein

expression studies using immunohistochemistry showed that NPC2 is highly expressed in

ovarian cancer tissue and absent in normal ovarian surface epithelium. In summary, clusterin

and NPC2 appear to play a role in ovarian cancer pathobiology and their role in EOC need to be

studied further.

iv

DEDICATION

I want to dedicate this PhD dissertation to my beloved parents, Benadict and Padmini

Gunawardana. They have always been the symbol of fortitude in my life and I hope I made

them proud.

I also dedicate this work to those who are suffering from cancer. I pray that this work benefits

them soon.

v

ACKNOWLEDGEMENTS

I am grateful to my supervisor, Dr. Eleftherios P. Diamandis, for allowing me to work in his

first-class laboratory. This thesis would not have been possible without his guidance and

support. Dr. Diamandis is a great mentor and I am honoured to have worked under his

supervision.

I like to acknowledge the following members of my PhD advisory and oral examination

committee for their guidance and advice during the last five years:

Dr. Sylvia Asa

Dr. Alexander Romaschin

Dr. Joe Minta

Dr. Margaret Fahnestock

I want to extend my thanks to the Department of Laboratory Medicine and Pathobiology at the

University of Toronto, as well as the Samuel Lunenfeld Institute and the Department of

Pathology and Laboratory Medicine at the Mount Sinai Hospital. In addition, I want to thank

Dr. Constantina Petraki for her help in this study and Dr. Peter Lobel, who was gracious to share

his research with us.

To the staff I had the honour and privilege of working with over the last five years at the

Advanced Centre for Detection of Cancer (ACDC) Laboratory, I thank you for your support. I

want to write a special thank you to the lab managers, Antoninus Soosaipillai, Tammy Earle,

vi

and Linda Grass. It is their hard work that keeps everyday lab operations running smoothly. I

like to thank you so much for your technical support, expertise, and of course, friendship.

vii

TABLE OF CONTENTS

ABSTRACT .................................................................................................................... II

DEDICATION.................................................................................................................IV

ACKNOWLEDGEMENTS...............................................................................................V

TABLE OF CONTENTS ...............................................................................................VII

LIST OF TABLES ...........................................................................................................X

LIST OF FIGURES ........................................................................................................XI

LIST OF ABBREVIATIONS........................................................................................ XIV

CHAPTER 1: GENERAL INTRODUCTION ..................................................................1

1.1 Ovarian Cancer..................................................................................................................................................... 21.1.1 Anatomy of the Human Ovary........................................................................................................................ 21.1.2 Ovarian Cancer ............................................................................................................................................... 3

1.4 Cancer Biomarkers............................................................................................................................................. 171.4.1 Types of Biomarkers..................................................................................................................................... 171.4.2 The Ideal Tumor Marker............................................................................................................................... 191.4.3 Cancer Antigen-125...................................................................................................................................... 20

1.5 Mechanisms of biomarker elevation in biological fluids................................................................................. 22

1.6 Proteomics and ovarian cancer ......................................................................................................................... 231.6.1 Principles and instrumentation...................................................................................................................... 231.6.2 Ovarian cancer proteomics: sources to mine for biomarkers ....................................................................... 25

1.7 Purpose and aims of the present study ............................................................................................................. 271.7.1 Rationale ....................................................................................................................................................... 271.7.2 Hypothesis .................................................................................................................................................... 281.7.3 Objectives ..................................................................................................................................................... 28

CHAPTER 2: PROTEOMIC ANALYSIS OF CELL-CULTURE SUPERNATANTS BY 2D-LC MASS SPECTROMETRY .................................................................................30

2.1 Introduction ........................................................................................................................................................ 31

2.2 Materials and Methods ...................................................................................................................................... 342.2.1 Cell lines ....................................................................................................................................................... 34

viii

2.2.2 Cell Culture................................................................................................................................................... 342.2.3 Sample Preparation ....................................................................................................................................... 342.2.4 Strong Cation Exchange Chromatography ................................................................................................... 352.2.5 Mass Spectrometry ....................................................................................................................................... 352.2.6 Data Analysis ................................................................................................................................................ 36

2.3 Results.................................................................................................................................................................. 382.3.1 Optimization of culture conditions ............................................................................................................... 382.3.2 Identification of Proteins by Mass Spectrometry ......................................................................................... 392.3.3 Identification of internal control proteins ..................................................................................................... 392.3.4 Intracellular and intercellular overlap........................................................................................................... 412.3.5 Cellular localization ...................................................................................................................................... 41

2.4 Discussion ............................................................................................................................................................ 48

CHAPTER 3: CANDIDATE SELECTION AND VERIFICATION IN SERUM BY ELISA......................................................................................................................................54

3.1 Introduction..................................................................................................................................................... 55

3.2 Materials and Methods ...................................................................................................................................... 563.2.1 Immunoassays............................................................................................................................................... 563.2.2 Biotinylation of detection antibody .............................................................................................................. 573.2.3 Clinical Specimens ....................................................................................................................................... 583.2.4 Statistical Analysis........................................................................................................................................ 58

3.3 Results.................................................................................................................................................................. 593.3.1 Selection of candidates ................................................................................................................................. 593.3.2 Construction of immunoassays ..................................................................................................................... 643.3.3 Preclinical Validation of candidates ............................................................................................................. 65

3.4 Discussion ............................................................................................................................................................ 72

CHAPTER 4: STUDY OF CANDIDATE PROTEIN EXPRESSION IN OVARIAN CANCER TISSUE BY IMMUNOHISTOCHEMISTRY...................................................75

4.1 Introduction ........................................................................................................................................................ 76

4.2 Materials and Methods ...................................................................................................................................... 824.2.1 Materials ....................................................................................................................................................... 824.2.2 Tumor specimens.......................................................................................................................................... 824.2.3 Immunostaining ............................................................................................................................................ 824.2.4 Evaluation of immunohistochemical staining............................................................................................... 83

4.3 Results.................................................................................................................................................................. 844.3.1 ADAM15 expression .................................................................................................................................... 844.3.2 Clusterin expression...................................................................................................................................... 854.3.3 EPCR Expression.......................................................................................................................................... 924.3.4 ICAM 5 Expression ...................................................................................................................................... 964.3.5 IGFBP5 Expression .................................................................................................................................... 1004.3.6 IGFBP7 Expression .................................................................................................................................... 1004.3.7 Integrin β4 Expression ................................................................................................................................ 104

ix

4.4 Discussion .......................................................................................................................................................... 111

CHAPETER 5: ANTIBODY PRODUCTION AND IMMUNOASSAY DEVELOPMENT FOR NPC2 ..................................................................................................................117

5.1 Introduction ...................................................................................................................................................... 118

5.2 Materials and Methods .................................................................................................................................... 1205.2.1 Biological specimens .................................................................................................................................. 1205.2.2 NPC2 purification from seminal plasma .................................................................................................... 1205.2.3 In-Gel Digestion ......................................................................................................................................... 1215.2.4 Mass spectrometric analysis ....................................................................................................................... 1225.2.5 PIM assay design ........................................................................................................................................ 1235.2.6 Rabbit immunization................................................................................................................................... 1245.2.7 Antibody purification.................................................................................................................................. 1245.2.8 Western Blotting ......................................................................................................................................... 1255.2.9 Biotinylation of detection antibody ............................................................................................................ 1255.2.10 Immunoassays........................................................................................................................................... 125

5.3 Results................................................................................................................................................................ 1285.3.1 Analysis of Cell Culture Supernatants........................................................................................................ 1285.3.2 Analyzing complex fluids for NPC2 .......................................................................................................... 1285.3.2 Development of Product Ion Monitoring Assay for NPC2. ....................................................................... 1285.3.4 Development of a high-throughput screening assay for NPC2. ................................................................. 1385.3.5 Development of Polyclonal anti-NPC2 antibody ....................................................................................... 1455.3.6 Construction of NPC2 immunoassay.......................................................................................................... 1495.3.5 Measuring NPC2 in serum.......................................................................................................................... 152

5.4 Discussion .......................................................................................................................................................... 160

CHAPTER 6: SUMMARY AND FUTURE DIRECTIONS ..........................................163

6.1 Summary ........................................................................................................................................................... 1646.1.1 Key Findings............................................................................................................................................... 1646.1.2 Proof of Hypothesis .................................................................................................................................... 167

6.2 Future Directions .............................................................................................................................................. 170

REFERENCES............................................................................................................172

x

List of Tables

Table Title Page

Table 1.1: Types of Malignant Ovarian Tumors 8

Table 1.2: Risk and Protective Factors in Ovarian Cancer 10

Table 1.3: International Federation of Obstetrics and Gynecology Staging 14

Table 2.1: Previously studied proteins in EOC that were identified in this study. 47

Table 3.1: List of 51 protein candidates. 60-61

Table 4.1: Properties of the IGFBP 1-7 78

Table 4.2: ADAM15 expression in ovarian tumors (proportion of positive cases) 86

Table 4.3: Clusterin expression in ovarian tumors (proportion of positive cases) 89

Table 4.4: EPCR expression in ovarian tumors (proportion of positive cases) 93

Table 4.5: ICAM5 expression in ovarian tumors (proportion of positive cases) 97

Table 4.6: IGFBP5 expression in ovarian tumors (proportion of positive cases) 101

Table 4.7: IGFBP7 expression in ovarian tumors (proportion of positive cases) 105

Table 4.8: Integrin β4 expression in ovarian tumors (proportion of positive cases) 108

Table 5.1: Number of peptides of NPC2 identified in ovarian cancer cell lines 129

Table 5.2: Tryptic peptides of NPC2 identified by mass spectrometry 132

Table 5.3: NPC2 expression in ovarian tumors (proportion of positive cases) 156

xi

List of Figures

Figure Title Page

Figure 2.1: Experimental workflow 40

Figure 2.2: Overlap of proteins identified in the three replicates for each cell line 43

Figure 2.3: Intercellular overlap of all proteins identified in this study 44

Figure 2.4: The number of proteins identified in each subcellular compartment for each cell line

45

Figure 2.5: The number of proteins identified in each subcellular compartment 46

Figure 2.6: Comparing the proteins identified in this study with those found in other proteomic profiling studies for ovarian cancer

53

Figure 3.1: The major biological functions associated with the 51 candidate proteins 62

Figure 3.2: The major diseases associated with the 51 candidate proteins 63

Figure 3.3. Initial screening results of the 8 candidates tested in serum of EOC patients and healthy individuals

67

Figure 3.4: Proteins that interact with clusterin 70

Figure 3.5: Proteins that interact with IGFBP6 71

Figure 4.1: Immunohistochemical expression of ADAM15 in the four major types of epithelial ovarian cancer

87

Figure 4.2: Immunohistochemical expression of ADAM15 in normal surface epithelium 88

Figure 4.3: Immunohistochemical expression of Clusterin in the four major types of epithelial ovarian cancer

90

Figure 4.4: Immunohistochemical expression of Clusterin in normal surface epithelium 91

Figure 4.5: Immunohistochemical expression of EPCR in the four major types of epithelial ovarian cancer

94

Figure 4.6: Immunohistochemical expression of EPCR in normal surface epithelium 95

Figure 4.7: Immunohistochemical expression of ICAM5 in the four major types of epithelial ovarian cancer

98

Figure 4.8: Immunohistochemical expression of ICAM5 in normal surface epithelium 99

xii

Figure 4.9: Immunohistochemical expression of IGFBP5 in the four major types of epithelial ovarian cancer

102

Figure 4.10: Immunohistochemical expression of IGFBP5 in normal surface epithelium 103

Figure 4.11: Immunohistochemical expression of IGFBP7 in the four major types of epithelial ovarian cancer

106

Figures 4.12: Immunohistochemical expression of IGFBP7 in normal surface epithelium 107

Figure 4.13: Immunohistochemical expression of Integrin β4 in the four major types of epithelial ovarian cancer

109

Figures 4.14: Immunohistochemical expression of Integrin β4 in normal surface epithelium

110

Figure 5.1: LC-MS/MS analysis of semi-purified NPC2 protein showing the identification of proteotypic peptide A (upper panel) and the MS/MS daughter ions produced after fragmentation of peptide A (lower panel)

133

Figure 5.2: LC-MS/MS analysis of semi-purified NPC2 protein showing the identification of proteotypic peptide B (upper panel) and the MS/MS daughter ions produced after fragmentation of peptide B (lower panel)

134

Figure 5.3: Direct ELISA for NPC2 in gel filtration fractions 136

Figure 5.4: PIM assay using the NPC2 standard for calibrating the mass spectrometer 137

Figure 5.5: PIM assay on malignant ovarian ascites sample 1 140

Figure 5.6: PIM assay on malignant ovarian ascites sample 2 141

Figure 5.7: PIM assay on malignant ovarian ascites sample 3 142

Figure 5.8: PIM assay on malignant ovarian ascites sample 4 143

Figure 5.9: PIM assay on malignant ovarian ascites sample 5 144

Figure 5.10: Ion-exchange fractions separated by SDS-PAGE 147

Figure 5.11: Determining the presence of anti-NPC2 antibodies in rabbit antisera 148

Figure 5.12: Verifying the specificity of the new rabbit polyclonal anti-NPC2 antibody 150

Figure 5.13: NPC2 calibration curve for the NPC2 sandwich-type ELISA 151

Figure 5.14: Levels of NPC2 in sera from patients with ovarian cancer and healthy individuals

154

Figure 5.15: Immunohistochemical expression of NPC2 in the four major types of epithelial ovarian cancer

157

xiii

Figures 5.16: Immunohistochemical expression of NPC2 in normal surface epithelium 158

Figure 6.1: Flow chart representing the criteria used for candidate selection 165

xiv

LIST OF ABBREVIATIONS

ACN, Acetonitrile

ALP, alkaline phosphatase

BSA, bovine serum albumin

CA125, carbohydrate antigen 125

CM, Conditioned media

CT, computed axial tomography

DFP, diflunisal phosphate

DNA, deoxyribonucleic acid

DTT, dithiothreitol

EDTA, ethylenediamine tetra-acetic acid

ELISA, enzyme-linked immunosorbent assay

EOC, Epithelial ovarian cancer

EPCR, Endothelial protein C receptor ESI, electrospray ionization

FBS, Fetal bovine serum

FIGO, International Federation of Gynecology and Oncology

xv

FT-ICR, fourier-transform ion-cyclotron resonance

GO, gene ontology

HNPCC, hereditary nonpolyposis colorectal cancer

IGFBP, Insulin-like growth factor binding protein IPI, international protein index

kDa, kilodalton

KLK6, kallikrein 6

m/z, mass-to-charge ratio

MALDI, matrix-assisted laser desorption/ionization

MRM, multiple reaction monitoring

MS, mass spectrometry

MS/MS, tandem mass spectrometry

P value, probability value

PIM, product ion monitoring

RNA, ribonucleic acid

SCX, strong cation exchange chromatography

SRM, single reaction monitoring

xvi

TOF, time of flight

VEGF, vascular endothelial growth factor

WHO, world health organization

1

Chapter 1: General Introduction

2

1.1 Ovarian Cancer

1.1.1 Anatomy of the Human Ovary

The ovaries are the female equivalent of the testes in males. They are a pair of nodular

bodies, each one situated on the left and right side of the uterus, in relation to the lateral pelvic

wall. The ovaries are responsible for generating ova (eggs) and the female sex hormones,

estrogen and progesterone. They are suspended by peritoneal folds and ligaments on either side

of the uterus and attached to the back of the broad ligament of the uterus, behind and below the

uterine tubes. On average they measure between 3 to 5 cm in length and weigh between 2 to 4

grams.

The surface epithelium, the stromal cells, and the oocytes are the main types of cells

found in the ovary. The surface epithelium, which is derived from the coelomic epithelium,

lines the external surface of the ovaries and is a single cell layer of flat-to-cuboidal cells. The

coelomic epithelium also gives rise to the peritoneum, endometrium, endocervix and

endosalpinx. The stroma is made up of soft tissue that has an abundant supply of blood vessels.

It consists, for the most part, of spindle-shaped cells with a minority of connective tissue

dispersed here and there. Due to the similarities in resemblance some anatomists liken stromal

cells to unstriped muscle, whereas others have described them as being similar to connective-

tissue cells. On the surface of the organ (below the surface epithelium) the stromal tissue is

more condensed, and forms a layer (tunica albuginea) composed of short connective-tissue

fibres, with fusiform cells amongst them. Interstitial cells resembling those found in the testis

may also be found in the ovarian stroma. Germ cells or oocytes are found near the periphery of

the ovaries. The granulose cells surround the germinal cells that form the follicles. The stroma

3

immediately surrounding the follicles differentiates into elongated cells known as theca cells,

which produce androgens such as androstenedione when stimulated by luteinizing hormone.

1.1.2 Ovarian Cancer

1.1.2.1 Symptoms of ovarian cancer

Clinically, ovarian cancers often present as a mass in the pelvis. Over 80% of patients

have symptoms even when the disease is limited to the ovaries (58). Symptoms for ovarian

cancer can be grouped into abdominal, gastrointestinal, genitourinary and pelvic response

symptoms (131). These symptoms include increased abdominal size, abdominal bloating,

fatigue, abdominal pain, indigestion, urinary frequency, pelvic pain, constipation, urinary

incontinence, back pain, pain with intercourse, early satiety, weight loss, nausea, bleeding with

intercourse, deep venous thrombosis and diarrhea (107). Given that these symptoms mimic

benign conditions, they are not suitable for early diagnosis. According to the most recent report

released by the Canadian Cancer Society there will be 2500 new cases of ovarian cancer in

2009. Approximately, 1750 will succumb to the disease in 2009. Furthermore, the report

predicts a 1.4% chance for a woman to develop ovarian cancer in her lifetime and a 1.1%

chance of dying from the illness. Regardless of the stage, the 5-year survival rate is

approximately 45%, however if caught in the early stages the 5-year survival rate jumps to 95%.

1.1.2.2 Types of ovarian cancer

Ovarian cancer is a heterogeneous disease and tumors can be categorized based on the

cells of origin. The majority of tumors of the ovaries fall into one of three major categories:

surface epithelial tumors, sex cord-stromal tumors and germ cell tumors. Approximately, 10-

15% of ovarian cancer cases are sex cord-stromal tumors. More than 50% of stromal cell

4

tumors are seen in postmenopausal women over the age of 50. Some cases are also seen in

young girls. Female reproductive hormones are produced in some stromal cell tumors. This

results in vaginal bleeding in postmenopausal women, or precocious puberty in young girls. In

rare situations, male hormones can be produced resulting in menstrual irregularities, hirsutism,

and virilism. Common types of malignant stromal cell tumors include granulosa cell tumors,

theca cell tumors, sertoli-leydig cell tumors, and hilar cell tumors. About 5-10% of ovarian

cancer cases are germ cell tumors, which arise from the oocytes. These are more common in

adolescent girls and accounts for approximately 60% of ovarian tumors in women below 20

years of age (123). Teratomas, dysterminomas, endodermal sinus tumors, and choriocarcinomas

are the typical examples of germ cell tumors.

Epithelial ovarian cancer (EOC), the most lethal among all ovarian malignancies, arises

from the ovarian surface epithelium and makes up 80% of all ovarian cancer cases (7, 69).

Since epithelial ovarian cancer accounts for the majority of ovarian cancers, research is mainly

focused on diagnosis and treatment of EOC. The present study also focuses on this particular

form of ovarian cancer.

EOC can be either benign or malignant. The benign tumors seldom spread from the

ovaries and are not associated with serious disease. Benign tumors include serous adenomas,

mucinous adenomas, and Brenner tumors (173). Malignant tumors of the ovarian surface

epithelium are known as carcinomas. These malignancies have the potential to spread into the

proximal and distal areas of the body and therefore can cause life-threatening disease. Based

on tissue morphology, EOC can be subdivided into four major types: Serous, mucinous,

endometrioid, and clear-cell carcinomas. In addition there are other minor types of EOC such as

malignant Brenner tumors and undifferentiated carcinomas (153).

5

Serous carcinomas of the ovary resemble those of the epithelium of the Fallopian tube. It

makes up 40-60% of the EOC cases and is the most aggressive histological type. Less than a

quarter of the cases are detected in the early stages (stage I and II). High-grade serous

carcinoma involves the surface of the ovary (often bilaterally), and the peritoneal membranes

with rapid carcinomatosis. Serous carcinomas have a broad spectrum of histological

appearances. The morphological heterogeneity of serous carcinomas may be a reflection of the

genetic heterogeneity. Most serous carcinomas show papillary and micropapillary architecture

with solid areas mixed in with chamber-like open spaces. Cytologically, serous carcinomas

typically contain columnar cells, but polygonal eosinophilic cells, clear cells, signet ring cells,

and spindle cells also exist. In some cases it is difficult to differentiate glandular or cribriform

serous carcinomas from endometrioid carcinomas (discussed later); microcystic serous

carcinomas from mucinous (discussed later); and clear cell containing serous carcinomas from

clear-cell carcinomas. Other features characteristic of serous carcinomas include the expression

of WT1 (73), p53 overexpression and p53 mutations (high-grade carcinomas) (98, 162), and loss

of BRCA1 expression in high grade tumors (5).

Endometrioid tumors are the second most common type of EOC and make up 10-20% of

EOC cases. These tumors resemble their endometrial counterparts. Tissue patterns containing

tubules, cribriform structures, solid, sheet-like growth, and papillae will be present in the

context of an endometrial tissue-like background. Given their similarity to endometrial tissue,

most endometrioid tumors are associated with endometriosis, endometrioid borderline tumors,

or coexisting tumors of the endometrium (15, 143). Most endometrial carcinomas have either

squamous or mucinous differentiation. In addition, these carcinomas may show secretory

features(165) and can also demonstrate sex cord-like features or spindle cells (178). Molecular

6

features that are characteristic of endometrioid carcinomas include the nuclear expression of the

estrogen receptor (ER), the progesterone receptor (PR), and β-catenin (55, 154). In addition the

genes encoding β-catenin, PI3K, and PTEN have been reported to have mutations in ovarian

endometrial carcinomas (25, 124, 129).

Mucinous tumors make up less than 3% of all epithelial ovarian cancer cases.

Indentifying intracytoplasmic mucin is mandatory for diagnosis, however, obvious mucin

expression can be absent in large parts of the tumor. They are formed by cells that often

resemble the intestinal epithelium and sometimes the endocervical epithelium. Similar to serous

papillary tumors, malignant mucinous tumors may contain papillary projections within cyst

cavities, large solid areas and areas of necrosis and haemorrhage. Most mucinous carcinomas

show transitions from intestinal mucinous borderline tumors to carcinomas. Mucinous

carcinomas express CK7 over CK20(169), lack expression of estrogen receptors(168), and lack

mesothelin and fascin expression (27). Finally, K-ras mutations are very common in mucinous

carcinomas (46, 75).

Clear-cell tumors make up 5-10% of malignancies of the ovary. Cells with hobnail

configurations (round expansion of clear cytoplasm with a narrow basal section containing the

nucleus) are often found in these tumors. Being predominantly solid or cystic, they often

contain polyp-like masses that protrude into the lumen. Clear cell tumors are typically

malignant and are the most lethal among all EOC histological subtypes. At least half of clear

cell tumors are associated with endometriosis. In fact, this type of carcinoma is difficult to

differentiate from serous and endometrioid carcinomas of the ovary. Immunohistochemically

speaking, clear cell carcinomas show low ER, PR, WT1, p53, and mib-1 expression (154).

7

Mutations in K-ras (127) and PTEN (146) have also been reported. However, solid markers that

can positively identify clear-cell ovarian carcinoma have not been found.

Other types of EOC include undifferentiated tumors, which consist of highly malignant

epithelial tumors that lack any specific differentiation with diffuse solid areas as the

predominant component. They do not resemble any of the above subtypes and tend to grow and

spread the fastest. Borderline tumors make up a large percentage of EOC cases (up to 10%).

They do not appear cancerous and are characterized as having low malignant potential.

Borderline tumors mostly behave as benign tumors and have good prognosis, but some may

recur after surgical removal and others may metastasize within the abdominal cavity. Most

borderline tumors are similar to serous and mucinous histologically and occur mainly in young

women. The list of malignant ovarian tumors is shown in Table 1.1.

8

Table 1.1: Types of Malignant Ovarian Tumors (173) 1. Common epithelial tumors

A. Serous tumors B. Mucinous tumors C. Endometrioid tumors D. Clear cell carcinomas E. Brenner Tumors F. Mixed epithelial tumors G. Undifferentiated carcinomas H. Unclassified tumors

2. Specialized stromal cell cancers A. Granulosa cell tumors B. Theca cell tumors C. Sertoli-Leydig cell tumors D. Hilar cell tumors

3. Germ cell tumors A. Teratomas B. Mature teratomas C. Immature teratomas D. Struma ovarli E. Carcinoids F. Dysgerminomas G. Embryonal cell carcinomas H. Endodermal sinus tumors I. Primary choriocarcinomas J. Gonadoblastomas

4. Soft tissue tumors not specific to the ovary 5. Unclassified tumors 6. Secondary (metastatic) tumors 7. Tumor-like conditions

9

1.1.3.1 Epidemiology

Accounting for approximately 3% of all new cancer cases in 2009, ovarian cancer ranks

fifth in cancer-related deaths in women. Although rare in comparison to breast cancer (1 in 7)

and prostate (1 in 6), ovarian cancer (1 in 59) is the most lethal gynecological malignancy and

overall one of the most lethal types of cancer, accounting for more deaths than endometrial and

cervical cancer combined (81). The median age of patients with epithelial ovarian cancer is 60

years (26). Among the risk factors for ovarian cancer, a strong family history of either ovarian

or breast cancer remains the most important one. However, the majority of ovarian cancer cases

are sporadic and only 5% of affected women have an identifiable genetic predisposition. Most

familial cases of EOC are related to mutations in the BRCA1/BRCA2 genes (26). Hereditary

ovarian cancer generally occurs 10 years earlier in comparison to familiar ovarian cancer cases

and is characterized by a trend towards an earlier age of diagnosis at each successive generation

(59). Ovarian cancer is a component of three hereditary cancer syndromes, namely breast-

ovarian cancer syndrome, site-specific ovarian cancer syndrome, and hereditary nonpolyposis

colorectal cancer (HNPCC)(59, 107). In addition to hereditary predisposition, other identified

risk factors include endocrine, environmental, dietary, and genetic factors (Table 1.2)(107).

More specifically, advancing age, nulliparity, hormonal therapy and exposure to talc have been

associated with increased risk (131). In contrast, the use of oral contraceptives, pregnancy, and

lactation are associated with reduced risk. Some of these findings suggest that continued

stimulation of the ovarian epithelium due to uninterrupted ovulation might increase the risk of

malignant transformation (172).

10

Table 1.2: Risk and Protective Factors in Ovarian Cancer (32, 107)

Risk Factors

Genetic Breast ovarian cancer syndrome, HNPCC, familial site-specific ovarian cancer syndrome

Environmental Exposure to talc or asbestos, smoking, genital deodorant, highest rates in industrialized countries of North America and Scandinavia, lowest rate in Japan, being Caucasian, being of Jewish descent

Dietary High consumption of meat and animal fat

Endocrine Nulliparity, early menarche, late menopause, use of infertility drugs, hormonal therapy, increases with the number of ovulatory events

Protective Factors

Dietary Consumption of vegetables, low-fat milk, lactose

Endocrine Oral contraceptives, pregnancy, tubal ligation, oophorectomy, hysterectomy, breast feeding

11

1.1.3.2 Etiology

There have been several hypotheses proposed regarding the beginnings of EOC. The

incessant ovulation hypothesis was developed when it appeared that women with a greater

number of ovulatory cycles have an increased risk of ovarian cancer (50). According to this

hypothesis, uninterrupted ovulation leads to a continuous cycle of damage and repair of the

surface epithelium. The repair mechanisms place the cells at an increased risk of developing

mutations and subsequent progression into a cancer. Additionally, higher ovulatory activity is

associated with more inclusion cysts and other changes in the ovarian surface, such as

invaginations. These inclusion cysts may be a suitable environment for ovarian cancer

development (51). Consistent with this hypothesis, women with multiple pregnancies,

increased time of lactation, and oral contraceptive use have a lower incidence of ovarian

cancer(64, 119, 172). However, this theory is weakened by the fact that progesterone-based oral

contraceptives that do not inhibit ovulation are equally effective as ovulation inhibiting

contraceptives (137). In addition, women with polycystic ovarian syndrome whose ovulatory

cycles are reduced, have a high risk of developing EOC (147).

Failure of the incessant ovulation hypothesis to explain certain observations such those

mentioned above led to the gonadotropin hypothesis. The pituitary gonadotropin hypothesis

suggests that increases in gonadotropins that initiate ovulation, persisting in high levels for years

following menopause are capable of stimulating the ovarian surface epithelial cells and inducing

malignant formation (34, 122, 128). In addition, gonadotropins are able to stimulate an

ovulation-like loss of the ovarian surface epithelial basement membrane (140). Since

inflammation is a well-known precursor to cancer development, the chronic inflammatory

processes of the ovarian surface epithelium may be a mechanism by which gonadotropin

12

stimulation and ovulation contribute to ovarian cancer risk (1, 122). Ovulation is an

inflammatory-like process involving multiple cytokines and proteolytic enzymes, and their

actions ultimately lead to tissue rupture. Inflammation may also provide an explanation for the

increased risk associated with talc and asbestos exposure, endometriosis, pelvic inflammatory

disease, and mumps infection (128).

The most recent theory hypothesizes that ovarian cancer does not begin in the ovary, but

at the distal fallopian tube. This hypothesis is supported by the fact that the majority of early

serous malignancies, detected in risk-reducing bilateral salpingo-oophorectomies (BSO) in

healthy women, were found in the distal fallopian tube and not the ovary. In addition, analysis

of mutations in TP53 in early serous malignancies of the distal fallopian tube and adjacent bulky

carcinomas of the ovary showed shared mutations (98). Though this theory may explain the

origin of serous carcinomas of the ovary, it does not explain endometrioid, mucinous, or clear

cell forms of ovarian cancer.

1.1.3.4 FIGO Classification

Ovarian cancer is classified based on the stage of the disease under the guidelines

established by the International Federation of Gynecology and Oncology (FIGO). This is

determined by considering the size of the tumor, the extent of tumor invasion into other tissues,

compromise of the lymphatic system, and establishment of distal metastasis. Typically, ovarian

cancer is classified into 4 stages: I, II, III and IV, where the first three stages are further

subdivided (Table 1.3) (66). Stage I tumors are those limited to one or both ovaries with the

tumor extending to the surface of the ovary by the third substage. Stage II identifies tumors

with pelvic extensions beyond the surface of the ovaries to the uterus, fallopian tubes and/or

other pelvic tissues by a pattern of spread called direct extension. Stage II tumors present with

13

ruptured capsules by the third substage. In stage III, the tumor extends intraperitoneally to

distant organs within the abdominal cavity and in stage IV, the tumor cells enter the circulation

and travel through the lymphatic system or the haematogenous circulation and metastasize to

lymph nodes and other organs in the body, including the pleural space, and the hepatic or

splenic parenchyma (36, 141).

The stage at diagnosis is established by thorough examination of the tumor and surgical

determination of disease progression (26). Patients diagnosed with FIGO stages I and II have a

5-year survival rate that exceeds 90% with surgery alone. However, those diagnosed with FIGO

states III and IV have significantly lower 5-year survival rates of 10-30% (21). Unfortunately,

only about 19% of cases are identified in the early stages (21, 81, 173). Patient prognosis and

treatment will vary depending on the stage of ovarian cancer.

1.1.3.5 Ovarian Cancer Screening, Detection, Treatment, Prognosis and Management

Unfortunately, only 19% of ovarian cancers are diagnosed at early stages (stage I or II)

while the tumor is still localized or confined to the ovary. About 7% are diagnosed with regional

(pelvic) spread and the vast majority (68%) are diagnosed with distant spread (abdomen and

extra-abdominal) (81). The fact that most cases are diagnosed at late stages is the major cause of

the high death rate of patients with ovarian cancer.

Screening is not suitable with ovarian cancer because no test or series of tests have been

found to be sufficiently sensitive or specific. Despite such limitations, current screening

methods for ovarian cancer consist of a combination of pelvic examinations, measurement of

serum cancer antigen-125 (CA 125) levels, and transvaginal or pelvic ultrasonography.

14

Table 1.3: International Federation of Obstetrics and Gynecology Staging (107) Stage Description

Stage I Stage IA Growth limited to one ovary, no ascites; no tumor on the external surface with capsules intact Stage IB Growth limited to both ovaries, no ascites; no tumor on the external surface with capsules

intact Stage IC Tumor stage IA or IB, but with tumor on the surface of one or both ovaries, with capsules

ruptured, malignant cells within ascites, and/or with positive peritoneal washings Stage II

Stage IIA Growth involving one or both ovaries with pelvic extension and/or metastases to the uterus and/or fallopian tubes

Stage IIB Growth involving one or both ovaries with advanced pelvic extension Stage IIC Tumor either stage IIA or IIB with tumor on the surface of one or both ovaries, with capsules

ruptured, malignant cells within ascites, and/or with positive peritoneal washings Stage III

Stage IIIA Tumor grossly limited to the true pelvis with negative nodes but with histologically confirmed microscopic seeding of abdominal peritoneal surfaces

Stage IIIB Tumor of one or both ovaries with histologically confirmed implants of abdominal peritoneal surfaces, none > 2cm diameter with negative nodes

Stage IIIC Abdominal implants > 2cm in diameter and/or positive retroperitoneal or inguinal nodes Stage IV Growth involving one or both ovaries with distant metastases. The presence of pleural effusion

with positive cytology allots a case to stage IV. The presence of parenchymal liver metastases allots a case to stage IV

15

These are normally performed annually or semi-annually, particularly in women with strong

family history of ovarian cancer (58, 81, 128).

Upon suspicion of ovarian cancer (based on symptoms and physical pelvic

examination), the levels of serum CA125 are measured along with transvaginal and abdominal

ultrasonography. In addition, a computed axial tomography (CT) scan of the abdomen and

pelvis is performed. Once diagnosed, exploratory laparotomy may lead to the resection of one

or both ovaries, fallopian tubes and/or uterus in addition to sampling of lymph nodes, liver and

suspicious sites within the abdomen to check for metastasis. The surgery provides a definite

diagnosis, identifies the histology and stage of the tumor, and removes the majority of the tumor

(26, 108). The surgeon will aim to remove as much tumor as possible in a process called optimal

debulking or cytoreduction, optimally leaving tumors no larger than 1cm. Ultimately, surgery is

performed to improve the patient’s response to chemotherapy (61, 108).

Additional treatment after surgery is dependent on the stage of the disease. Well-

differentiated stages IA or IB ovarian cancer that has been surgically removed usually requires

no further treatment. Patients with stage IA or IB ovarian cancer with a poorly differentiated

tumor and stages IC, II, III and IV disease are classified as high risk where chemotherapy

treatment is required. The optimal regimen for postoperative chemotherapy to eradicate residual

disease is currently still being studied, but combination therapy with a platinum compound such

as cisplatin or carboplatin, or a taxane/platinum combination such as paclitaxel/carboplatin are

given. The paclitaxel and platinum combination achieves clinical response in approximately

80% of patients (26, 108).

16

Although disease stage is the key prognostic factor, other factors such as volume of

cancerous tissue remaining after surgery, histological type, age of patient (over 69 fare worse),

and patient’s overall condition or performance status, may also influence the outcome. FIGO

staging is currently used to determine 5 year survival rates where women with stage I have a 5-

year survival rate of 80% to 95%, whereas women with stage III have only 10% to 30% chance

of surviving five years (21). Histologic type and grade are also significant, with clear-cell,

mucinous, and poorly differentiated tumors being the worst prognostically (107).

Although treatment strategies have improved ovarian cancer management, overall

survival rate has not improved. This is largely due to diagnosis at late stages where 5-year

survival rates are low. Early stage ovarian cancer often evades the current screening

procedures, as patients tend to be asymptomatic or exhibit common symptoms of upper

abdominal disease (12). Thus, nearly 80% of patients present with late stage disease when the

cancer has already metastasized to distant organs and the prognosis is poor.

17

1.4 Cancer Biomarkers

1.4.1 Types of Biomarkers

Serological biomarkers are an effective and relatively non-invasive approach for the

early detection, diagnosis, prognosis and management of many types of diseases, including

ovarian cancer. A biomarker is defined as a quantifiable characteristic that is objectively

measured and evaluated as an indicator of a normal biologic process, a pathogenic process, or a

pharmacologic response to a therapeutic intervention. Typically, they are endogenous molecules

that can be measured in bodily fluids or tissues with the ability to distinguish between disease

and normal states.

Cancer biomarkers may appear in different types and forms, including DNA, mRNA,

proteins, metabolites, or processes such as apoptosis, angiogenesis or proliferation (65).

Additionally, different functional subgroups of proteins, such as enzymes, glycoproteins,

oncofetal antigens and receptors, may serve as useful biomarkers. Furthermore, tumor changes

such as genetic mutations, amplifications, translocations and changes in microarray profiles

(signatures) may also be utilized as tumor markers.

Tumor markers may be detected in a variety of fluids, tissues and cell lines as they are

often produced by the tumor itself or by other tissues in response to the presence of cancer or

other associated conditions, such as inflammation. By measuring the levels of such markers

through a serological test, tumor markers can be used for population screening, differential

diagnosis in symptomatic patients, and for clinical staging of cancer. In addition, they may also

be used to estimate tumor volume, to evaluate response to treatment, and to assess recurrence

through monitoring or as prognostic indicators for disease progression (76).

18

Biomarkers are classified as being diagnostic, prognostic, or predictive. Diagnostic

markers are applied in disease detection and in identifying a given type of cancer in an

individual. To minimize false positive and false negatives rates, diagnostic markers are expected

to have high sensitivity and specificity. Screening markers are a specific type of diagnostic

markers where they are used to examine the general population for a disease (48). Currently,

there is no perfect screening marker for ovarian cancer.

Prognostic markers, on the other hand, are used once the disease state has been

established and are applied to determine the probability of a patient responding to therapy in

order to improve the accuracy of medical prediction and the etiology of the disease following

tumor resection. These markers are expected to predict the likely course of the disease, its

recurrence, and influence the type of therapy provided to the patient. Currently, FIGO staging is

the major prognostic factor for ovarian cancer to identify patient prognosis and treatment for

ovarian cancer patients.

Lastly, predictive markers are used to predict the response to a drug before treatment is

initiated. Optimally, it is able to classify individuals as likely responders or non-responders to a

particular treatment. Often, predictive markers arise from array-type experiments that make it

possible to predict clinical outcome from the molecular characteristics of the patient’s tumor.

Unfortunately, other than definitive diagnosis by biopsy and histopathology, there is currently

no single diagnostic, prognostic, or predictive tumor marker with acceptable sensitivity and

specificity for ovarian cancer.

19

1.4.2 The Ideal Tumor Marker

The World Health Organization (WHO) lists specific criteria that a biomarker must

satisfy. According to WHO, a good screening test must meet the following criteria (31):

1. There are significant mortality statistics for the target disease and occurrence in

the population

2. Disease progression should be well characterized

3. Early stage treatment of the disease should offer improved outcome

4. Public acceptance of the screening test

5. Availability of effective treatment options for individuals with advanced disease

6. Suitable treatment and diagnostic facilities

7. Policy outlining who can be subjected to treatment

8. Cost-effective screening

9. High positive predictive value, negative predictive value, sensitivity and

specificity

In addition to the above criteria, an ideal tumor marker should be measured easily,

quickly, reliably and cost-effectively using an assay with high analytical sensitivity and

specificity (48). Its differential expression in a significant portion of the patient population

should be characteristic of the cancer of interest, and rarely occur for other conditions or for

normal patients (41).

The ideal marker should be produced by the tumor cells and enter the circulation in order

for it to be detected by a non-invasive serological test. The marker should be present at low

levels in serum of healthy or benign disease patients and increase significantly in cancer

(preferably in one cancer type). Optimally, an ideal marker is present in detectable (or higher

20

than normal) quantities at early or preclinical stages and the quantitative levels of the tumor

marker should reflect the tumor burden. The assay for this marker should demonstrate high

diagnostic sensitivity (low false negatives) and high specificity (low false positives).

Current tumor markers for ovarian cancer, such as carbohydrate antigen 125 (CA125),

suffer from low diagnostic sensitivity and specificity when used alone. Consequently, they are

used in conjunction with imaging, biopsy and associated clinicopathological information prior to

setting a diagnosis or prognosis.

1.4.3 Cancer Antigen-125

Currently, the clinically accepted serum marker for ovarian cancer is carbohydrate

antigen 125 (CA125), a high molecular weight mucin (glycoprotein) with unknown function

(177). It was discovered initially by a radioimmunoassay in patients with advanced ovarian

cancer (9). It is expressed by fetal amniotic and coelomic epithelium, and in adult tissues that

are derived from the coelomic (mesothelial cells of the pleura, pericardium and peritoneum) and

Mullerian (tubal, endometrial and endocervial) epithelia. Normal epithelium of the ovaries does

not express CA125 on the surface (83).

While CA125 may be the best ovarian cancer biomarker discovered to date, its utility as

a screening marker is limited due to its high false positive rates. It is elevated in other

malignancies such as uterine, fallopian, colon and gastric cancer (77, 166) as well as in 1% of

the normal population, particularly in non-malignant conditions such as pregnancy,

menstruation and endometriosis (9-11, 77, 166). In addition, many prospective studies of

screening have revealed major limitations of CA125 also in its sensitivity (16). Specifically, the

sensitivity of CA125 is more than 90% for women with advanced stage ovarian cancer, but the

21

sensitivity for stage I ovarian cancer decreases to approximately 50%. As a result, its clinical

use for the early detection of ovarian cancer is limited (16, 77, 78). Therefore, CA125 is neither

sensitive nor specific enough to be used as a diagnostic biomarker.

Serum CA125 levels greater than 35 U/mL are considered elevated (166). These levels

may occur one to two years prior to conventional diagnosis (11, 44, 45, 166, 182). Screening

using CA125 may detect a proportion of ovarian cancer cases before symptoms arise (68, 78,

79). CA125 has been found to be particularly useful in detecting early relapse (112).

CA125 has also been shown to play a significant role in prognosis. Some studies have

demonstrated that concentrations of CA125 in the serum decrease with tumor regression, and

increase with progression in 74 to 95% of cases (13, 166). In addition, CA125 has been able to

predict survival outcomes in women with CA125 levels greater than 65 U/mL (116). Although

CA125 may be promising in its prognostic value, the current major prognostic factor is the

FIGO stage. Other conventional prognostic markers include variables such as tumor grade, size,

histological subtype, residual tumor after surgery, and patient age. However, ovarian cancer is a

highly heterogeneous disease, thus, cancers with similar clinical profiles may have different

outcomes.

Although CA-125 is used currently in clinical settings for diagnosis and prognosis, its

limitations in sensitivity and specificity warrant the development and identification of novel

ovarian cancer biomarkers that would complement CA125 in facilitating early disease detection,

determination of prognosis, and the development of more individualized and efficient treatment

plans for ovarian cancer patients.

22

1.5 Mechanisms of biomarker elevation in biological fluids

Protein levels are physiologically maintained in biological fluids. In disease states,

proteins may become elevated as a result of the disease by several mechanisms. These include

and are not limited to gene over-expression; angiogenesis, invasion and destruction of tissue

architecture; and finally increased protein secretion and shedding.

First, increased protein quantities may be due to increases in the specific gene or

chromosome copy number (gene amplification), epigenetic modifications such as DNA

methylation, and increased transcriptional activity. Increased transcriptional activity is often due

to the imbalance between gene repressors and activators.

Second, tissue invasion by the tumor may allow direct release of molecules into the

interstitial fluid, reabsorbed by the lymphatics and subsequently into the blood. In the case of

epithelial cancer types, proteins must break through the basement membrane of the invading

tumor before entering the circulation.

Third, as 20-25% of all proteins are secreted, elevated protein levels may occur due to

aberrant secretion or shedding of membrane-bound proteins containing an extracellular domain

(ECD). In addition, single nucleotide polymorphisms may cause alterations in the signal peptide

of proteins resulting in atypical secretion patterns (80). Cancer-associated glycoproteins may be

released into the circulation due to the change in the polarity of the cancer cells. Also, increased

protease expression may lead to increased ECD cleavage of membrane bound proteins resulting

in increased circulating levels of these cleaved products.

23

1.6 Proteomics and ovarian cancer

Proteomics focuses on the large-scale determination of gene and cellular function

directly at the protein level. The field of proteomics is a collection of various technical

disciplines, all of which contribute to protein analysis. One powerful proteomic approach

focuses on de novo analysis of proteins or protein populations isolated from cells or tissues.

Studies of cellular proteomes are challenging due to the high degree of complexity and the low

abundance of many proteins requiring highly sensitive analytical techniques in order to identify

these proteins. Among proteomic techniques, mass spectrometry (MS) has become the main

method used in the analysis of complex protein samples. It has an unparalleled ability to acquire

high-content quantitative information about biological samples of enormous complexity and

subsequently to use this data to identify proteins with high sensitivity and specificity.

1.6.1 Principles and instrumentation

MS identifies proteins by measuring the mass and charge of individual molecules and

atoms with high detection sensitivity and molecular specificity. This process is carried out in the

gas phase on ionized analytes, as the motion of gaseous molecules can be manipulated (156).

Typically, mass spectrometers consist of an ion source, a mass analyzer that measures the mass-

to-charge ratio (m/z) of the ionized analytes, and a detector that registers the number of ions at

each m/z value (35).

In order to volatize and ionize the proteins or peptides out of a solution prior to mass

spectrometry analysis, electrospray ionization (ESI) and matrix-assisted laser

desorption/ionization (MALDI) are the most commonly used ionization sources. ESI

encompasses three different processes: Droplet formation, droplet shrinkage and gaseous ion

24

formation. ESI-based systems ionize the analytes out of a solution as it is passed through an

electrostatic field (3-4 kV) generating a fine mist of charged droplets (156). Nitrogen gas is

often added to assist in evaporation of solvent from these charged droplets. ESI is often coupled

to liquid-based separation tools (chromatographic or electrophoretic). MALDI sublimates and

ionizes the samples out of a dry, crystalline matrix via a laser beam of short pulses (57). The

matrix absorbs energy at the wavelength of the laser and the energy is then transferred to the

samples as the laser beam causes evaporation of the matrix. Integrated liquid-chromatography

ESI-MS is normally used to analyze complex samples while MALDI-MS is used to analyze

relatively simple peptide mixtures.

Upon ionization, the peptide ions enter the first mass analyzer (MS1) which separates

gas-phase ions generated from the ionization source according to their m/z ratio. The molecular

mass of each peptide is determined in this step. Next, they are directed into a collision cell

where the peptides collide with neutral gas molecules and become fragmented. The m/z values

of the resultant fragments are measured in the second mass analyzer (MS2), producing a tandem

mass spectrum. This is carried out under high vacuum to prevent ions from colliding with other

species. The mass spectrum is then analyzed by various algorithms such as MASCOT,

SEQUEST and X!Tandem. The amino-acid sequence of each peptide is generated and matched

against the human genome sequence to identify possible proteins (105).

Ion motion in the mass analyzer can be manipulated by electric or magnetic fields in

order for ions to reach the detector in an m/z-dependent manner. Commonly used mass

analyzers include beam (time-of-flight (TOF) and quadrupole) and trapping (ion-trap and

Fourier-transform ion-cyclotron resonance (FT-ICR)) analyzers. Ion-trap mass analyzers store

25

and manipulate ions in time rather than in space. Quadrupole ion-trap instruments use an

oscillating electric field to “trap” and determine the masses of the ions.

1.6.2 Ovarian cancer proteomics: sources to mine for biomarkers

Potential biomarkers may be identified in various sources such as tumor tissues and biological

fluids such as serum, plasma, disease associated fluid, and cancer cell lines. These sources may

then be analyzed using mass spectrometry in order to identify the proteins (174).

In regards to ovarian cancer, the serum or plasma of ovarian cancer patients may be

compared to the serum or plasma of healthy controls. This biological fluid is an optimal source

to mine for biomarkers, as secreted proteins of the cancer should be found in the circulation. In

addition, if the biomarker is detectable within the serum of patients and controls, serological

tests to measure biomarker levels in plasma and serum are relatively non-invasive and

inexpensive. As the blood contains more than 100,000 different protein forms with abundances

spanning over 10-12 orders of magnitude (4), biomarkers are most likely present in this fluid.

Unfortunately, the search for tumor-derived biomarkers within this fluid is challenging as 20 of

the most abundant plasma proteins (concentration ranges in the mg/mL range) account for 99%

of the total protein mass and impedes the detection of lower abundance tumor antigens by mass

spectrometry (4). Potential tumor markers are expected to exist in the low nanogram to

picogram per millilitre concentration range. However, the presence of highly abundant proteins

such as albumin and immunoglobulins suppresses the ionization of low abundance proteins.

Currently, up-front fractionation techniques are performed in order to remove major proteins in

the blood in order to detect these potential low abundant tumor markers.

26

Alternatively, tissue samples from the disease may be another source to mine for

potential biomarkers, such as comparing normal ovarian tissue against ovarian tumors (174).

Hypothetically, certain proteins originating from the tissue could subsequently appear and be

monitored in the blood stream. The shedding and secretion of tumor proteins into the

bloodstream are expected to occur due to leaky capillary beds, protease cleavage and high rates

of cell death within the tumor mass. However, these samples are often complex incorporating

many different types of cells. Often, tumor biopsies may not simply contain tumor tissues but

also include blood components as well as normal tissue. Thus, proteomic analysis of tumor

biopsies may also identify proteins from circulating cells, normal tissues and from plasma thus

only identifying a small population of tumor related proteins (82).

27

1.7 Purpose and aims of the present study

1.7.1 Rationale

Proteins are the biological effector molecules in the body. Given that they are more

dynamic than DNA or RNA, and they reflect the physiological status of the human body,

proteins seem the best suited for biomarker research. With respect to cancer, classical tumor

markers such as carcinoembryonic antigen (CEA) and alpha-feto protein (AFP) were discovered

in the ‘60s with the development of novel and relatively sensitive immunological techniques

such as radial immuno-diffusion. The assays for the ovarian cancer marker, CA-125, were

developed in the late 70’s and early 80’s with the introduction of the monoclonal antibody

technology. The latest FDA approved biomarker for ovarian cancer, HE4, was discovered with

the use of DNA microarray technology. Based on these historical facts, the discovery of novel

biomarkers is intimately connected with technological advancement. With mass spectrometry

being the latest technology introduced to biomarker research, it is reasonable to predict that new

markers will be discovered using this technology

Many cancer biomarkers can be discovered using mass spectrometry as the tool for

discovery. These molecules, however, are difficult to identify because their concentrations in

serum and/or biological fluids are too low and therefore cannot be measured or purified, unless

specific immunological reagents and highly sensitive ELISA methods are available. As

discussed before, the complexity of serum is a problem for biomarker research using mass

spectrometry. Thus, in order to identify novel cancer biomarkers within the initial discovery

phase, a less complex sample needs to used.

28

1.7.2 Hypothesis

Tumors secrete or shed proteins, and these proteins have the potential to enter

circulation. Indeed, the best ovarian cancer biomarkers such as CA-125 and HE4 are shed and

secreted respectively by ovarian tumors and are found in the circulation. It is reasonable to

assume that cell lines derived from ovarian tumors secrete or shed proteins that are similar to the

tumor of origin. Given that conditioned media of ovarian cancer cell lines are relatively less

complex than serum, mining conditioned media avoids the drawbacks of serum proteomics

while providing useful clues to ovarian cancer biology.

We hypothesize that:

1. Proteins secreted or shed by ovarian cancer cell lines are similar to those secreted

or shed by primary ovarian tumours.

2. These proteins can be identified by two-dimensional liquid chromatography-

coupled mass spectrometry.

3. These proteins can be measured in biological fluids such as serum using antibody

based immunoassays and/or mass spectrometry-based single reaction

monitoring/multiple reaction monitoring assays.

4. Some proteins may serve as biomarkers for early detection or prognosis of

ovarian cancer.

1.7.3 Objectives

1. Utilize emerging proteomic technologies such as mass spectrometry to develop a

biomarker discovery platform.

29

2. Demonstrate the feasibility of using cell line models for biomarker discovery.

3. Demonstrate the power in using an integrated proteomic approach (combining

data from several proteomic studies) for selecting candidates for further study.

4. Construct antibodies and develop sensitive immunoassays for candidates that do

not have commercial immunological reagents available.

30

Chapter 2: Proteomic Analysis of Cell-Culture Supernatants by

2D-LC Mass Spectrometry

Reproduced with permission from The Journal of Proteome Research. Comprehensive analysis of conditioned media from ovarian cancer cell lines identifies novel candidate markers of epithelial ovarian cancer. Gunawardana CG, Kuk C, Smith CR, Batruch I, Soosaipillai A, Diamandis EP. J Proteome Res. 2009 Oct;8(10):4705-13 Copyright 2009 American Chemical Society

31

2.1 Introduction

Ovarian cancer (OvCa) kills more women than any other gynaecological malignancy.

For a cancer that accounts for only 3% of new cases, it is the 5th largest killer. The reason for

the high case-to-fatality rate is that it is often diagnosed when the cancer has metastasized to

other organs. The 5-year survival rate for patients with advanced disease (stage III & IV) is 10-

30% (26). In contrast, the 5-year survival rate for patients diagnosed with early-stage disease

can be as high as 94% (26). These numbers clearly support the need for early diagnosis.

In general, ovarian malignancies arise in 3 major cell types. Epithelial ovarian cancer

(EOC) accounts for 80% of the cases and is found on the surface epithelium. Stromal cell

tumors arise in the connective tissue below the surface epithelium and account for 10% of cases.

The third type arises from germ cells and accounts for less than 10% of cases. This study

focuses on EOC, and in particular the serous, endometrioid, clear-cell and mucinous histological

types.

The clinically accepted biomarker for EOC is CA-125 (8). Approximately 85% of

clinically advanced ovarian carcinomas can be identified by measuring CA-125 levels(10, 44).

However, this molecule is a poor marker for early detection due to frequent false positive and

false negative results (114). Other markers that have shown some clinical relevance in EOC are

HE4 (40), osteopontin (86), the carbohydrate antigens CA 15-3 and CA 19-9 (56), inhibin (139),

and several members of the kallikrein family (kallikreins 5, 6, 8,10, 11 and 14) (23, 102, 111,

148, 150). None of these proteins, however, have been effective early-detection biomarkers nor

have they reached the clinical efficacy of CA-125 for detecting recurrence and monitoring

therapy.

32

Many strategies exist to uncover novel biomarkers for cancer, including gene expression

profiling, protein microarrays, gene translocation/fusion analysis, peptidomics, and mass

spectrometry (MS)-based profiling (95). MS-based proteomic studies using EOC tissue (17,

181), ascites fluid (17, 60, 93, 181), and cancer cell lines (49, 53) have contributed greatly to the

list of potential protein markers. However, the selection and validation of these candidate

biomarkers have been major rate-limiting steps.

In the late ‘70’s the development of a novel technology, namely the monoclonal

antibody, helped in discovery of many tumor markers including CA-125, CA15-3, and PSA(9).

Therefore, it is conceivable that with new emerging technologies such as the mass spectrometer,

novel tumor markers can be identified. The assumption that suitable cancer biomarkers for

diagnosis and prognosis will be either secreted or plasma membrane proteins is reasonable given

that:

1. Secreted proteins are more likely to enter circulation

2. Membrane proteins have the potential to be cleaved and therefore can enter

circulation

3. These proteins can be measured using robust immunological techniques

4. All currently known biomarkers (e.g. PSA, CA-125, and CA 15-3) are secreted or

shed proteins.

Given that serum is too complex for mass-spectrometric based discovery projects, a less

complex sample is essential. Therefore, we examined the proteome of cell-culture supernatants

from ovarian cancer cell lines. In this study, we report a shotgun proteomics approach to

33

analyze the conditioned media of the HTB-75, TOV-112D, TOV21G, and RMUG-S cell lines.

Each cell line represents the serous, endometrioid, clear-cell, and mucinous EOC histological

types, respectively.

34

2.2 Materials and Methods

2.2.1 Cell lines

HTB-75, TOV-112D, and TOV-21G cell lines were purchased from the American Type

Culture Collection (ATCC), Manassas, VA. The RMUG-S cell line was purchased from the

Japanese Collection of Research Bioresources. (Osaka, Japan). HTB-75 cells were maintained

in RPMI medium containing 10% fetal bovine serum (FBS). TOV-112D and TOV 21G cell

lines were grown in a 1:1 mixture of MCDB 105 medium and Medium 199, containing 10%

fetal bovine serum. RMUG-S cells were maintained in Ham’s F12 medium containing 10%

fetal bovine serum. All media for cell culture was purchased from Invitrogen Canada Inc.

(Burlington, Ontario, Canada).

2.2.2 Cell Culture

Each cell type was seeded in T-175 cm2 cell culture flasks and cultured to 80%

confluency in normal growth medium (2 days). Eight flasks were grown per cell line and cells

were washed 3 times with 30 ml of phosphate buffered saline (PBS). Following the washes, 30

ml of chemically defined serum-free CDCHO medium (Invitrogen) supplemented with 8 mM

glutamine (Invitrogen) was added to each flask. HTB-75, TOV-112D, and TOV-21G cell lines

were grown for 48 hours, whereas RMUG-S was grown for 72 hours in serum-free CDCHO

medium. Following the growth in serum-free medium (SFM), the conditioned media (CM)

were collected and centrifuged to remove cellular debris.

2.2.3 Sample Preparation

A total of 240 ml of CM were collected per cell line (8 flasks, each having 30 ml of

CM). CM in the eight flasks (of each cell line) were first combined, centrifuged to remove

35

cellular debris and then separated into four aliquots (60 ml) per cell line. Each aliquot

represented a technical replicate, and thus 4 replicates were available per cell line. In this study,

we processed 3 replicates per cell line. Each replicate was dialyzed (3.5 kDa molecular mass

cut-off) against 5 litres of 1 mM ammonium bicarbonate with 2 buffer exchanges at 4oC.

Following dialysis, the replicates were lyophilized. Each lyophilized replicate was denatured

using 8M urea, reduced with 13mM dithiothreitol (DTT, Sigma), and then alkylated using 500

mM iodoacetamide (Sigma). Following reduction and alkylation, the replicates were desalted

using NAP5 columns (GE Healthcare). Each replicate was lyophilized and then trypsin-

digested (Promega) overnight at 37oC. Following trypsin digestion, each replicate was

lyophilized once more.

2.2.4 Strong Cation Exchange Chromatography

Each trypsin-digested and lyophilized replicate was resuspended in 120 µl of mobile

phase A [0.26 M formic acid in 10% acetonitrile(ACN)]. The sample was injected into a

PolySULFOETHYL ATM column with a 200-Å pore size and diameter of 5 µm (The Nest

Group, Inc.) containing a hydrophilic, anionic polymer (poly-2-sulfoethyl aspartamide). A 1-hr

separation was performed on an HPLC system (Agilent 1100) using a mobile phase B

containing 0.26 M formic acid in 10% acetonitrile and 1M ammonium formate. The eluate was

monitored at a wavelength of 280 nm. Fractions were collected every 5 minutes after the start

of the run at a flow rate of 200 µl/min.

2.2.5 Mass Spectrometry

Fractions 6-11 obtained from strong cation exchange (SCX) chromatography were used

for mass spectrometric analysis. Each fraction was loaded onto a ZipTip C18 pipette tip

36

(Millipore; catalogue number ZTC18S096) and eluted in 4µl of Buffer B [90% ACN, 0.1%

formic acid, 10% water, 0.02% Trifluoroacetic acid (TFA)]. The eluate was mixed with 80 µl

of Buffer A, and 40 µl were injected via an autosampler into an Agilent 1100 series HPLC.

The peptides were first injected onto a 2-cm C18 trap column (inner diameter, 200 µm), and

then eluted from the trap column into a resolving 5-cm analytical C18 column (inner diameter,

75 µm) with an 8 µm tip (New Objective). The LC setup was coupled online to a 2-D linear ion

trap (LTQ, Thermo Inc.) mass spectrometer using a nano-ESI source in data-dependent mode.

Each fraction was run on a 120 min gradient. The eluted peptides were subjected to MS/MS.

DTAs were created using the Mascot Daemon (version 2.16) and extract_msn. We used the

following parameters for DTA creation: minimum mass, 300 Da; maximum mass, 4000 Da;

automatic precursor charge selection; minimum peaks, 10 per MS/MS scan for acquisition; and

minimum scans per group, 1.

2.2.6 Data Analysis

Mass spectra from each fraction were analyzed using Mascot (Matrix Science, London,

UK; version 2.2) and X!Tandem (Global Proteome Machine Manager, version 2006.06.01)

search engines on the non-redundant International Protein Index (IPI) human database version

3.27 (containing 67528 entries). Up to one missed cleavage was allowed, and searches were

performed with fixed carbamidomethylation of cysteines and variable oxidation of methionine

residues. A fragment tolerance of 0.4 Da and a parent tolerance of 3.0 Da were used for both

Mascot and X!Tandem, with trypsin as the digestion enzyme. Six DAT files (Mascot) and six

XML files (X!Tandem) were generated per replicate, per cell line. The DAT and XML files

were uploaded and analyzed using Scaffold (v01_05_19, Proteome Software Inc., Portland,

OR). Peptide identifications and protein identifications were accepted if they could be

37

established with greater than 95% probability using the PeptideProphet algorithm and greater

than 80% probability using the ProteinProphet algorithm, respectively. The number of

identified peptides was set to at least one. All biological samples were searched using the

MudPIT (multidimensional protein identification technology) option. Sample reports were

exported from Scaffold and the identified proteins were assigned a cellular localization based on

information available from Swiss-Prot, Genome Ontology (GO), and other publicly available

databases. To calculate the false positive error rate, each fraction was analyzed using a

“sequence-reversed” decoy IPI human database version 3.27 by Mascot and X!Tandem and data

analysis was performed as mentioned above.

38

2.3 Results

2.3.1 Optimization of culture conditions

To optimize the growth conditions of cell lines, the following was performed:

1. For good coverage of the proteome of supernatants, at least 1 mg of total protein was

required. Therefore cell lines were grown such that at least 1 mg of total protein can

be obtained after growth in serum-free medium (SFM)

2. Cells were grown such that contamination from intracellular proteins was minimized.

3. Recovery of secreted proteins and shed proteins (membrane origin) was maximized

in cell culture supernatants while considering factors 1 & 2.

Each cell type was seeded in T-175 cm2 cell culture flasks and cultured to 80%

confluency in normal growth medium. Flasks were washed three times with 30 ml of phosphate

buffered saline (PBS). Following the washes, 30 ml of chemically defined serum-free CDCHO

medium (Invitrogen) supplemented with 8 mM glutamine (Invitrogen) was added to each flask.

LDH levels were measured every day for 5 days. KLK5 and KLK6 levels were also measured

every day for 5 days to monitor protein secretion.

LDH levels showed a steady increase in each of the cell lines and then a quick rise

starting around day 3. Total protein on day 2 for HTB-75, TOV-112D, and TOV21G met the

1mg requirement (data not shown). RMUG-S cells were grown for an extra day so that 1 mg of

total protein could be recovered. KLK5 and KLK6 levels showed a steady increase in HTB75

and RMUG-S cell lines over time. The levels of KLK5 and KLK6 in TOV112D and TOV21G

were below the detection limit of the ELISA. Based on these observations, we selected 3 days

39

of growth for the RMUG-S cell line, and 2 days of growth in serum- free medium for the HTB-

75, TOV-112D, and TOV21G cell lines.

2.3.2 Identification of Proteins by Mass Spectrometry

The workflow and experimental design are illustrated in (Figure 2.1). Approximately 29

proteins were found in the negative controls for TOV-112D and TOV-21G; 82 proteins for

HTB-75; and 45 proteins for RMUG-S. Proteins found in the negative controls were compared

with the list of proteins of their counterpart cell lines. We eliminated proteins that were

common to both the negative controls and the conditioned media only if the total spectra of a

given protein in the negative controls were greater than 10 % of the total spectra for the same

protein in conditioned media. After eliminating proteins found in the negative controls, 1208

proteins were found in HTB-75; 1252 proteins for TOV-112D; 885 for TOV-21G; and 467 for

RMUG-S. The proteins lists were combined and the redundancies were removed. In total,

2039 unique proteins were identified.

2.3.3 Identification of internal control proteins

One advantage of our approach to biomarker discovery was that one could measure the

levels of endogenous internal control proteins. We knew a priori that HTB75 and RMUG-S

produced KLK5 whereas the TOV-112D and TOV-21G cell lines did not. We measured KLK5

levels by ELISA in the conditioned media of all cell lines prior to mass spectrometric analysis.

We hypothesized that if our approach is valid, then the mass spectrometry should detect KLK5

protein in the supernatants of HTB75 and RMUG-S only.

40

Figure 2.1: Experimental workflow

41

Indeed, this was the case. KLK5 levels in HTB75 conditioned media were 3 ng/ml and in

RMUG-S, levels were greater than 10 ng/ml. The KLK5 levels in TOV-112D and TOV-21G

conditioned media were below the detection limit of the KLK5 ELISA.

2.3.4 Intracellular and intercellular overlap

We examined the overlap of proteins identified in the three replicates for each cell line

analyzed. The overlap was 70 % for HTB75; 63 % for TOV-112D; 40 % TOV-21G; and 66%

for RMUG (Figure 2.2). We suspect the lower reproducibility of the TOV-21G cell-line was

due to differences in sample handling. That is, the third replicate of TOV-21G supernatant was

analyzed using a new reverse-phase chromatographic column on the mass spectrometer and

therefore may have been a source of variability. Despite the low overlap for the TOV-21G cell

line, 75 % of the proteins in the TOV-21G conditioned media were found in at least two of the

replicates.

We next examined the intercellular overlap of proteins identified. Of the 2039 proteins,

155 proteins were common to all four cell-lines (Figure 2.3). However, approximately 31 % of

the proteins were unique to HTB75; 27 % to TOV-112D; 15 % to TOV-21G; and 25 % to

RMUG-S. These proteins may be OvCa subtype specific and may have the potential to be used

as a subtype specific marker. This was beyond the scope of the present study and therefore we

did not explore this. All proteins identified in this study are available as part of supplementary

data in Gunawardana et al.(63)

2.3.5 Cellular localization

Proteins were cross-referenced with the Gene Ontology and Swiss-Prot databases to

determine their subcellular localizations. A significant proportion of proteins were from

42

intracellular locations such as the cytoplasm, golgi, endoplasmic reticulum, and the nucleus.

This was most likely due to cell lysis, which is unavoidable with cultured cells. Figure 2.4

depicts the distribution of proteins based on subcellular location for each cell line. Figure 2.5

depicts the distribution of proteins based on subcellular location for all 2039 proteins identified.

Approximately a fifth of the proteins identified in this study were either extracellular or

plasma membrane proteins. Williams et al. recently published a comprehensive review listing

the proteins that have been studied as biomarkers in serum or ascites fluid in EOC (173). We

cross-referenced our list of extracellular and plasma membrane proteins with the

aforementioned list and the common proteins are listed in Table 2.1. Known markers of ovarian

cancer such as CA-125, CA 15-3, HE4, KLK6, and mesothelin were identified in this study.

43

Figure 2.2. Overlap of proteins identified in the three replicates for each cell line.

Three replicates per cell line (HTB75, TOV112D, TOV21G, and RMUG-S) were processed and

analyzed. For each cell line, the majority of proteins identified were found in all three

replicates.

44

Figure 2.3: Intercellular overlap of all proteins identified in this study. A total of 2039

proteins were identified.

45

Figure 2.4. The number of proteins identified in each subcellular compartment for each

cell line. All proteins were cross-referenced with the Gene Ontology database. The cytoplasmic

proteins included those classified as cytoskeletal by Gene Ontology. The organellar designation

includes proteins located in the mitochondria, endoplasmic reticulum, the Golgi, the nucleus,

peroxisomes, and lysosomes. Unclassified proteins are those that either did not have a Gene

Ontology classification or whose classifications were ambiguous and thus could not be placed in

the other four categories. There is redundancy in this data as some proteins were placed in

more than one compartment.

46

Figure 2.5. The number of proteins identified in each subcellular compartment. The 2039

proteins identified were cross-referenced with the Gene Ontology database to determine their

subcellular location. The cytoplasmic proteins included those classified as cytoskeletal by

Gene Ontology. The organellar designation includes proteins located in the mitochondria,

endoplasmic reticulum, the Golgi, the nucleus, peroxisomes, and lysosomes. Unclassified

proteins are those that either did not have a Gene Ontology classification or whose

classifications were ambiguous and thus could not be placed in the other four categories. There

is redundancy in this data as some proteins were placed in more than one compartment.

47

Table 2.1: Previously studied proteins in EOC that were identified in this study

Protein Subcellular location

Apolipoprotein A1 Extracellular

CA 125 Membrane

CA 15-3 Membrane

Cathepsin L Extracellular

Epidermal Growth Factor Receptor Membrane

Fibronectin Extracellular

Fibulin Extracellular

Human epididymal protein 4 Extracellular

Inhibin Extracellular

Interleukin-6 Extracellular

Kallikrein-6 Extracellular

Macrophage-colony stimulating factor Extracellular

Mesothelin Extracellular

Osteopontin Extracellular

α-1 antitrypsin

Extracellular

48

2.4 Discussion

In this study, a shotgun proteomics approach using LC-MS/MS was used to identify

proteins in conditioned media from 4 ovarian cancer cell lines. This technology has been

applied for biomarker discover previously (49, 94, 145). We hypothesized that the proteome of

the conditioned media of ovarian cancer cell lines may provide clues as to which proteins are

secreted by primary ovarian tumors. Mascot and X!Tandem databases were used to identify

over 2000 unique proteins. To our knowledge, this is one of the largest repositories of proteins

identified for ovarian cancer.

Serum is a fruitful source of potential markers for ovarian cancer. It contains more than

100,000 protein forms with concentrations ranging 10-12 orders of magnitude (4). The 20 most

abundant proteins make up 99 % of the total protein. The skewed protein distribution in serum

is a major challenge when MS-based strategies are used in pursuit of low-abundance cancer

biomarkers (37, 38). The main problems are:

1. Peptides from high abundance proteins outcompete their low abundance peptide

counterparts for ionization.

2. Peptides from the high abundance proteins suppress the ionization of the low

abundance peptides

On the grounds of these difficulties and more, we selected to analyze conditioned medium from

ovarian cancer cell lines, which is less complex than serum, yet is relevant to ovarian cancer

pathobiology. Furthermore, cell lines are easy to maintain and propagate, and offer an

inexhaustible source of mRNA and proteins. This material, in turn, can be rapidly processed

49

for profiling experiments using technologies such as DNA microarrays and mass spectrometry.

In addition, the biological variation between samples from the same cell line is low, thus

allowing greater reproducibility compared to tissue and serum samples. There are however,

several disadvantages of using cell culture supernatants for biomarker discovery. These are

listed below and are described in detail by Kulasingam et al.(96) :

1. A single cell line cannot model the heterogeneity of cancer.

2. There are multiple variants of the same cell line in circulation.

3. The interactions between the surrounding environment and the tumor that influence

its development are absent in a cell culture system.

4. Does not model the complex interactions among different cell types within the ovary.

5. Does not provide clues as to the causes of ovarian cancer.

In this study, we analyzed the cell culture supernatants of four cancer cell lines (focusing

on extracellular and plasma membrane proteins), each representing a histological type of

epithelial ovarian cancer. The HTB-75, TOV-112D, TOV-21G, and RMUG-S cells lines are

commonly used cell lines in ovarian cancer studies(85, 96, 133, 144). To represent serous

carcinoma we selected the HTB-75 cell line, the proteome of which is similar to tumor cells

originating from serous carcinoma of the ovary(49). The TOV-112D cell line represents the

endometrioid histological type. The proteome of this cell line clusters closer to cell lines

originating from endometrioid cancers of the ovary than to cell lines originating from other

histological types of EOC(171). The gene expression profile of TOV-21G, which represents

clear-cell carcinoma, is different from cell lines originating from other histological types (164).

Thus, we believe that HTB-75, TOV-112D, and TOV-21G provide a distinct look at ovarian

cancer. The choice of a mucinous carcinoma cell-line was RMUG-S. Studies comparing the

50

gene or protein expression profile of RMUG-S with mucinous carcinoma of the ovary were not

available in the literature. To our knowledge, this study is the first comparative proteomic study

conducted using this particular cell line.

A total of 2039 proteins were identified from the four cell lines. Of these, 420 proteins

were either extracellular or plasma membrane proteins. The proportion of extracellular and

membrane proteins (21%) relative to the total number proteins in this study is lower than studies

conducted on breast (34%)(94) and prostate cancer cell lines (39%)(145). We suspect that the

difference between the proportions found in this study and those of the other ones is due to the

following reasons:

1. Differences in bioinformatics. In this study, we attempted to improve upon the

searching strategies employed in the breast cancer and prostate cancer studies by not

only using gene ontology, but also employing the curated Swiss Prot database to

verify our extracellular and membrane protein classifications. By doing this we

noticed, for example, that some membrane proteins that were originally classified as

“plasma membrane” by gene ontology, were classified as membrane but of an

organellar origin. When there was a discrepancy between gene ontology and Swiss

Prot, we used the curated classifications of the Swiss Prot database.

2. We do not know whether or not a proteomic study on supernatants of cell lines of

different tissue origin will always provide the same proportion of extracellular and

plasma membrane proteins. It is possible that the ovarian cell lines we used had a

lower proportion of extracellular and plasma membrane proteins in the culture

media.

51

3. Sample preparation is a source of great variability, which may have contributed to

the discrepancy.

Furthermore, a large proportion of proteins (29%) with unknown Gene Ontology annotations

were identified and categorized as unclassified (Figure 2.4). Some of these proteins may indeed

be extracellular or plasma membrane proteins.

Of the 420 extracellular and membrane proteins identified, 94 were found in plasma by

HUPO(63, 126). The small overlap may be due to several reasons. First, some extracellular and

membrane proteins identified in this study have low abundance in plasma. With plasma being

very complex, it is reasonable to assume that mass spectrometry is unable to identify these

proteins. In addition, the elimination half-life for some proteins may be very short, meaning that

they are either removed from the circulation rapidly, or are eliminated within their

microenvironment before they can enter the circulation. Furthermore, some proteins may be

localized to particular compartments or microenvironments in the body and thus never enter the

circulation. Finally, some proteins are sensitive to sample handling and therefore are degraded

during the experiment.

The Kislinger group (60) and the Hanash group (49) recently published two major

proteomic studies using ascites fluid and ovarian cancer cell lines. Approximately, 44% of the

proteins identified in our study overlapped with the Hanash study whereas 29% of our proteins

overlapped with those found in the Kislinger study (Figure 2.5). However, comparing just the

extracellular and plasma membrane proteins, 75% of our proteins overlapped with those of the

Hanash study. Taking all three studies together, a repository of 8256 proteins can be

constructed, a valuable resource of proteins for further study in ovarian cancer. We have

contributed an additional 1091 proteins that were not identified in the Kislinger and Hanash

52

studies. A point to note is that only 555 proteins (7%) were common to all three studies. This is

most likely due to the differences in experimental approach, sample types, the inherent

variations in mass spectrometric analysis, and different bioinformatic platforms.

53

Figure 2.6: Comparing the proteins identified in this study with those found in other

proteomic profiling studies for ovarian cancer. The lists generated by Faca et al. (49) and

Gortzak-Uzan et al. (60) were compared with our list. A repository of 8256 proteins can be

generated.

54

Chapter 3: Candidate Selection and Verification in Serum by

ELISA

Reproduced with permission from The Journal of Proteome Research. Comprehensive analysis of conditioned media from ovarian cancer cell lines identifies novel candidate markers of epithelial ovarian cancer. Gunawardana CG, Kuk C, Smith CR, Batruch I, Soosaipillai A, Diamandis EP. J Proteome Res. 2009 Oct;8(10):4705-13 Copyright 2009 American Chemical Society

55

3.1 Introduction

Advances in MS-based proteomic technologies have helped create large datasets. These

enormous datasets have given rise to the hypothesis that new clinically relevant biomarkers will

be found within them. However, these optimistic predictions have yet to be fulfilled (136).

One of the greatest challenges in finding biomarkers is in selecting candidates for validation.

The critical question is what criteria to use when selecting candidates from a large list.

Considering the fact that the number of new proteins being discovered is increasing, selecting

the “best” candidates for testing proves more difficult. In fact, testing of candidates is limited

by several factors including the lack of suitable reagents, instrumentation, and manpower.

As mentioned in the preceding chapter, we discovered over 2000 proteins by analyzing

the conditioned media of four ovarian cancer cell lines. We hypothesize that novel biomarkers

can be found within our dataset. However, the criteria for selecting candidates are at the

investigators’ discretion. There are many strategies that can be employed to simplify a large

dataset including comparing proteomic lists with mRNA expression lists to select candidates

that are overexpressed in cancer, comparing several proteomic datasets to identify those proteins

that a consistently found in cancer, and focusing on a subset of proteins in a list (i.e. plasma

membrane proteins, proteins involved in adhesion, etc.) In this chapter we address the

aforementioned issue of selection criteria, and present a robust step-wise method to produce a

master list of candidates for ovarian cancer.

56

3.2 Materials and Methods

3.2.1 Immunoassays

IGFBP5, IGFBP6, βIG-H3, and cystatin C ELISA kits were purchased from R&D

Systems. The IGFBP4 kit was purchased from DSL Inc. and the clusterin assay was purchased

from ALPCO Laboratories. Immunoassays for IGFBP4, cystatin C, and clusterin were

performed according to the manufacturers instructions. Assays for IGFBP5, IGFBP6, and βIG-

H3 were also performed according to the manufacturers instructions but with a modification to

the detection step (see below). Some assays were not designed for use with serum and

therefore, required optimization (see Results section).

Non-biotinylated polyclonal and monoclonal antibodies to vasorin, EPCR, and IGFBP7,

were purchased from R&D Systems, as were the recombinant proteins used as protein

calibrators. Sandwich-type ELISAs were constructed in-house using a monoclonal or a

polyclonal antibody for antigen capture and a biotinylated polyclonal antibody for detection.

White polystyrene microtitre plates were coated with either 100ng/100µl (vasorin and IGFBP7)

or 200ng /100µl (EPCR) of monoclonal or polyclonal antibody in coating buffer (50mM Tris

buffer, 0.05% sodium azide, pH 7.8) and stored at room temperature overnight. Fifty

microlitres of protein calibrators or samples, and 50µl of assay buffer [50mM Tris, 6% BSA,

0.01% goat IgG, 0.1% bovine IgG (Sigma-Aldrich Inc, St. Louis MO), 0.005% mouse IgG

(Fortron Bio Science Inc, Morrisville, NC), 0.05% sodium azide, pH 7.8] with 0.5M KCl

(vasorin and IGFBP7) or without KCl (EPCR) were added to wells and incubated for 90 min

with shaking at room temperature. The plates were washed 6 times with washing buffer (5mM

Tris, 150mM NaCl, 0.05% Tween-20, pH 7.8). Approximately 100µl of biotinylated detection

57

antibody (125ng/ml in assay buffer containing 0.5M KCl) were added to each well and

incubated for 1 h at room temperature with shaking. The plates were then washed six times with

the washing buffer.

Detection of IGFBP5, IGFBP6, and βIG-H3 was modified from the manufacturer’s

instructions and performed the same way as that for vasorin, IGFBP7, and EPCR.

Approximately 100µl (5ng/well) of alkaline phosphatase-conjugated (ALP) streptavidin

(Jackson ImmunoResearch) in sample buffer (6% BSA, 50 mM Tris, 0.06% sodium azide, pH

7.8) was added to each well and incubated for 15 min with shaking at room temperature. The

plates were washed 6 times with the wash buffer, and then 100 µL of substrate buffer [0.1 mol/L

Tris buffer, pH 9.1, containing 0.5 mmol/L diflunisal phosphate (DFP), 0.1 mol/L NaCl, and 1

mmol/L MgCl2] were added to each well and incubated for 10 min with shaking at room

temperature. Approximately 100µl of developing solution (1 mol/L Tris base, 0.15 mol/L

NaOH, 2 mmol/L TbCl3, 3 mmol/L EDTA) were added to each well and incubated for 1 min

with shaking at room temperature. The fluorescence (615 nm) was measured with an

EnVision™ 2103 time-resolved fluorometer (Perkin Elmer).

3.2.2 Biotinylation of detection antibody

Biotinylated polyclonal antibodies to IGFBP5, IGFBP6 and βIG-H3 were provided with

their kits. Biotinylated polyclonal antibodies to vasorin, and EPCR were purchased from R&D

Systems. Approximately 50 ng of polyclonal anti-IGFBP7 antibody was incubated with 50 ng

of biotin in 0.5 M NaHC03 for 1 h. This was used as the detection antibody for the IGFBP7

assay. To verify the biotinylation reaction, sandwich-type ELISAs were first constructed (as

mentioned previously) using the biotinylated and the unbiotinylated versions of each antibody.

Purified recombinant proteins specific for each antibody were used as the standard. The

58

fluorescence (measured at 615 nm) was compared for the biotinylated and unbiotinylated

antibody.

3.2.3 Clinical Specimens

Serum samples (approximately 10 ml) were collected from stage III-IV EOC patients

and age-matched normal controls. Blood was initially collected in BD Vacutainer® SST™

tubes containing clot activator and serum separator gel. Tubes were inverted five times, allowed

to clot for 30 minutes, and centrifuged for 10 minutes at 1000-1300 x g in a swing bucket

centrifuge. The separated serum was then aliquoted and stored at -80 oC. Our protocols have

been approved by the Institutional Review Board at the University Health Network, Toronto,

Ontario, Canada.

3.2.4 Statistical Analysis

Assuming that cancer cases and normal cases are independent and their distributions are

non-parametric, the Mann-Whitney test was used to determine statistical significance when

comparing the concentrations of candidate biomarkers in normal and ovarian cancer sera.

59

3.3 Results

3.3.1 Selection of candidates

We used the following arbitrary criteria to pick candidates for analysis in serum of EOC cases

and healthy individuals:

1. The set of extracellular and membrane proteins was chosen as the starting point. It is

reasonable to hypothesize that extracellular (secreted) and membrane proteins (ones that

are shed) are more likely to enter the circulation. Comparing the list of the 228

extracellular and 192 plasma membrane proteins with that of the plasma proteome

published by HUPO (126), there was an overlap of 65 extracellular proteins and 29

plasma membrane proteins. We eliminated known high-abundance proteins with

concentrations greater than 5 µg/ml in plasma. Some proteins, such as clusterin, were

kept as candidates since their levels in patients with ovarian cancer have not been

reported in the primary literature.

2. Next, this set of extracellular and membrane proteins was compared with a list of 289

extracellular and membrane proteins of a separate study from our lab on the proteome of

ascites fluid (93). Seventy-two proteins overlapped and these were selected for further

investigation.

3. We further eliminated proteins that have been reported previously as serological markers

of ovarian cancer. By applying this criterion, 21 proteins were eliminated. The

remaining 51 proteins are listed in (Table 3.1). The major biological functions and

diseases associated with these proteins are illustrated in Figures 3.1 & 3.2.

60

Table 3.1: List of 51 protein candidates.

61

62

Figure 3.1: The major biological functions associated with the 51 candidate proteins. The

list of 51 proteins (see Table 3.1) was analyzed using Ingenuity Pathway Analysis. The top 10

biological functions associated with the 51 candidates are shown. The negative log of the P

value is shown on the y-axis. The greater the negative log of the P value, the greater the number

of proteins associated with a given function.

63

Figure 3.2: The major diseases associated with the 51 candidate proteins. The list of 51

proteins (Table 3.1) was analyzed using Ingenuity Pathway Analysis. The top 10 diseases

associated with the 51 candidates are shown. The negative log of the P value is shown on the

y-axis. The greater the negative log of the P value, the greater the number of proteins associated

with a given disease.

64

4. We searched for commercial ELISA kits for the 51 protein candidates. For candidates

that did not have a commercial ELISA, we searched for monoclonal or polyclonal

antibodies to construct in-house immunoassays.

5. Proteins that did not have commercial ELISA kits or antibodies were not studied further.

Of the remaining proteins that possessed an ELISA or antibodies, nine proteins were

selected for preliminary validation with patient sera. These proteins were cystatin c,

insulin-like growth factor binding protein 4, -5, -6, and -7, clusterin, vasorin, endothelial

protein C receptor (EPCR), and βIG-H3.

3.3.2 Construction of immunoassays

The immunoassays of the nine candidates were optimized before testing serum samples. We

used the following step-wise approach to construct the immunoassays:

1. The cystatin C and IGFBP4 ELISA kits were available in 96-well plate format with a

pre-coated capture antibody. The clusterin ELISA kit was a competitive binding assay

with pre-coated clusterin antigen. These assays had been already optimized for serum

studies and therefore they were performed according to the manufacturer’s instructions.

2. For the other analytes, recombinant proteins of each candidate were used as standards,

and assays were optimized to produce a linear standard curve. For some candidates,

both a monoclonal and a polyclonal antibody were available. For these, sandwich-type

immunoassays using both monoclonal-polyclonal and polyclonal-polyclonal antibody

configurations were constructed.

3. Each ELISA was next tested for its efficacy in detecting endogenous protein. We used

ascites fluid that was positive for each candidate as the test sample. Mass spectrometric

65

analysis verified the presence of each candidate in the ascites fluid, as described

elsewhere (93).

4. Lastly, we measured each analyte in serial dilutions of serum to examine the relationship

between the signal measured and the corresponding dilution. All assays, except

IGFBP5, produced linear dilution curves.

3.3.3 Preclinical Validation of candidates

Since we were unable to establish a workable IGFBP5 immunoassay, we could not

validate this candidate in this study. The eight remaining candidate proteins were evaluated

using sera from EOC cases (n=10) and normal healthy women (n=20). For 6 of the candidates

there was no significant difference between groups. A significant difference was seen

(p=0.0002, Mann-Whitney U test) between the EOC cases and healthy controls for clusterin,

with levels in EOC being higher (Figure 3.3). IGFBP6 was also significantly different

(p=0.002, Mann-Whitney U) between the EOC cases and healthy controls, with levels in EOC

being lower than the controls.

To examine whether the difference between normals and EOC cases for clusterin and

IGFBP6 is due to differences in gene expression, we searched the Oncomine gene expression

database (135) for DNA microarray data on these two proteins in ovarian cancer and healthy

tissue. Data showing clusterin mRNA expression in healthy and ovarian cancer tissue was not

available.

However, quantitative real-time PCR results presented by Hough et al. show that clusterin

mRNA is overexpressed in ovarian cancer tissue (72). Data for IGFBP6 showed that mRNA

expression is lower in serous ovarian cancer compared to normal ovarian tissue. To further

66

verify our findings, we conducted a search for immunohistochemistry data on both proteins

using the Human Protein Atlas (HPA) database (www.proteinatlas.org). A detailed description

of the HPA site is presented by Berglund et al(18). We searched for tissue data using the

following search parameters:

1. Moderate to strong staining in at least 3 patients with ovarian cancer

2. Negative staining in normal ovaries.

67

68

Figure 3.3. Initial screening results of the 8 candidates tested in serum of EOC patients

and healthy individuals. Normal designates women without ovarian cancer. OvCa designates

individual serum samples from EOC cases. Mann-Whitney test was used to calculate P values

and comparisons that are significantly different from each other (p < 0.01) are indicated with an

asterisk (*). Horizontal bar through each data set shows the median.

69

Clusterin showed staining in 5 out of 12 ovarian cancer tissue samples (data is not shown here

but is available publicly at the HPA). Normal ovarian tissue showed no staining, however data

was available for the stromal and follicular regions only. IGFBP6 did not pass our search

criteria as normal ovarian tissue (stromal tissue) showed weak staining (data available at the

HPA).

Furthermore, in order to elucidate the global cellular functions of these two proteins, we

examined clusterin and IGFBP6 using Ingenuity Pathway Analysis. The data showed clusterin

to be involved in several biological functions pertinent to tumor pathology including cell

development, growth and proliferation and movement. In addition, the major diseases associated

with clusterin were cancer, connective tissue disorders and endocrine disorders. The

interactome of clusterin is shown in Figure 3.4. The major biological functions for IGFBP6

were cell movement, growth and proliferation, and cell development. The major diseases

associated with IGFBP6 were cancer, skeletal and muscular disorders, and respiratory disease.

The interactome of IGFBP6 is shown in Figure 3.5.

70

Figure 3.4: Proteins that interact with clusterin. Molecular interactions associated with

clusterin (nexus node) were analyzed using Ingenuity Pathway Analysis. Solid lines show direct

interactions, whereas dotted lines show indirect interactions.

71

Figure 3.5: Proteins that interact with IGFBP6. Molecular interactions associated with

IGFBP6 were analyzed using Ingenuity Pathway Analysis. Solid lines represent direct

interactions, whereas dotted lines represent indirect interactions.

72

3.4 Discussion

Most investigations concerning biomarker research thus far have focused on analyzing

single types of specimens. That is, the research is limited to the use of only one source of

biomarkers such as cell lines, tissue specimens, or other biological fluids. In this study, we

used an integrated approach to mine for biomarkers. A key issue faced in this study was

applying reasonable criteria to choose meaningful candidates. The criteria are dependent on the

experimental questions being asked. In our analysis, we were interested in the extracellular and

membrane proteins since these are likely to enter the circulation and have a higher chance of

being measurable by a sensitive assay such as an ELISA. Therefore, our first criterion was to

select extracellular and membrane proteins only. However, a drawback of using cancer cell

lines is that they are no longer identical genetically or proteomically to the cancer from which

they originate. Therefore, candidates chosen exclusively from a list of proteins secreted or shed

by ovarian cancer cell lines may be biologically irrelevant to ovarian cancer.

Ascites fluid bathes the ovarian tumor and it is reasonable to assume that some proteins

found in ascites fluid originate from the tumor itself or its microenvironment. Therefore, by

selecting proteins that are common to both cell lines and ascites fluid, the list can be narrowed to

proteins that are biologically relevant to ovarian cancer. Indeed, by applying this criterion,

many well-documented markers of ovarian cancer were found in our study including HE4 (40,

67) and KLK6 (39, 70, 142) (see Table 2.1). CA-125 was also found in the conditioned media

of cell lines, but was not identified in the ascites study to which we compared our list of proteins

due to exclusion of proteins greater than 30 kDa (93). Altogether, our study demonstrates the

power of comparing the proteome of cell lines with that of a clinically relevant biological fluid

73

to identify new markers. This strategy is transferable to other cancers such as lung, pancreatic,

and liver.

Our final list of candidates included 51 proteins. Ingenuity pathway analysis revealed

that two of the major disease types associated with these proteins were cancer and reproduction.

This is encouraging given that our aim is to find biological markers of ovarian cancer. In

addition, some of the major molecular functions associated with these proteins include cell-to-

cell interaction, cellular function and maintenance, and cell growth and proliferation. These

functions are known to be important for tumor growth and metastasis (100).

A major bottleneck in identifying markers using a proteomic approach is candidate

validation. It is imperative that good antibodies and immunological assays are developed to

evaluate the numerous potential markers identified in studies so far. In our study, some

promising candidates could not be studied due to the lack of immunological reagents. We

analyzed 8 proteins that had ELISA kits or antibodies available. From this panel, both IGFBP6

and clusterin showed significant differences between EOC cases and healthy individuals.

Current standards imply that, a good biomarker is one that is preferably elevated in

tissues or biological fluids; clusterin showed such promise as a potential marker. To our

knowledge, clusterin levels in the serum of EOC cases have not been reported previously.

Regarding the candidates that did not show promise in serum, they cannot be dismissed since

their role in the pathogenesis of ovarian cancer needs to be determined. Although clusterin

showed promise, its effectiveness as a marker for early detection still remains an open question.

Our initial screen did not use serum from early stage EOC, and therefore, the use of clusterin to

detect early stage EOC cannot be ascertained. In addition, the sample size (n=30) in this study

74

is relatively low. Therefore further studies to test clusterin need a large cohort (n >100) with a

substantial number of early stage cancer patients.

Immunohistochemistry data available from the HPA show that clusterin is expressed in

greater amounts in ovarian cancer tissue relative to healthy. The immunohistochemistry data is

supported by the gene expression data published by Hough et al.(72) showing upregulated

clusterin mRNA in ovarian cancer. The results of our ELISA are concordant with the studies

mentioned above. These results raise the question as to why clusterin is upregulated in ovarian

cancer. Clusterin is important in several cellular functions including apoptosis (2, 30), cell

migration (99, 113), and cell development (149). Recent evidence suggests that clusterin may

be involved in ovarian cancer-related processes. Findings by Park et al. (130) showed that high

clusterin expressing ovarian cancer cells are resistant to Paclitaxel and that high clusterin

expression correlated with poor survival. Further studies are ongoing to understand the

pathobiological role of clusterin in ovarian cancer.

75

Chapter 4: Study of Candidate Protein Expression in Ovarian

Cancer Tissue by Immunohistochemistry

The immuohistochemical study was aided by our collaborating pathologist, Dr. Constantina Petraki, Evangelismos Hospital, Athens, Greece.

Dr. Petraki supervised the staining of tissues and provided expert interpretation of the data.

76

4.1 Introduction

Most proteins that enter circulation can be measured in body fluids such as blood, saliva,

and urine using highly sensitive technologies such as the ELISA. However, not all proteins

originating from tumors enter circulation. There may be several reasons for this:

1. The protein is intracellular.

2. The protein is cleared from the tumor microenvironment rapidly by proteases.

3. The protein degrades before it has a chance to enter circulation.

4. The protein is bound to other proteins that remain localized to the tumor

microenvironment.

Given the aforementioned reasons, if a tumor-specific protein is not found in the circulation, it

does not mean that it cannot be used as a biomarker. Such proteins can be used for tumor

staging, identifying histological type, prognosis, response to treatment, and so forth.

In the preceding chapter, we measured serum levels of 8 candidates, namely clusterin,

cystatin C, IGFBP4, -6, and -7, Vasorin, EPCR, and βIG-H3. Of these, clusterin and IGFBP6

showed a difference between cancer and normal sera. Very little protein expression data is

available regarding many of these candidates with respect to EOC. Therefore we looked at

protein expression using immunohistochemistry in four subtypes of EOC (serous, endometrioid,

clear-cell, and mucinous) and normal healthy ovarian surface epithelium. Of the 8 candidates

studied by ELISA, the antibodies to clusterin, IGFBP5, IGFBP7, and EPCR were also suitable

for immunohistochemistry. Therefore we looked at their protein expression in cancerous and

healthy tissue. In addition, we also studied the expression of three other proteins, namely

ADAM15, integrin β4, and ICAM5. These proteins were found in the proteomic study of cell

77

culture supernatants, and antibodies optimized for immunohistochemistry were available.

Therefore we included them as part of the study.

The ADAM (a disintegrin and metalloproteinase) proteins are members of the metzincin

superfamily of matrix metalloproteinases (22, 43). To date, 21 functional ADAMs have been

studied in humans (42). These proteins have two clearly defined biological functions:

Proteolysis and cell adhesion. Their protease activity is focused on transmembrane proteins

such as precursor forms of growth factors. The ADAMs have been implicated in shedding

ligands of HER family of receptor (42). The HER family of receptors has been implicated in the

progression and development of multiple cancer types (125). ADAM15 appears to be important

to metastasis, in that loss of ADAM15 decreased metastasis to the bone in a prostate cancer cell

line model. In addition, mice deficient in ADAM15 develop smaller tumors upon melanoma

cell injection.

The insulin growth factor binding proteins (IGFBPs) are a family of proteins that bind to

and modulate the activity of the IGF ligand. IGF signalling plays a major role in the growth,

differentiation, and proliferation of mammalian cells. IGF signalling has been implicated in

breast cancer biology, and thus the role played by IGFBPs in cancer is an exciting area of

research. There are seven IGFBPs, numbered 1-7. Table 4.1 summarizes the properties of these

seven IGFBPs. The activity of IGF is related to the level of circulating IGFBP.

78

Table 4.1: Properties of the IGFBP 1-7 (taken from Subramanian et al. (157))

IGFBP Production site a Molecular Weight (kDa)

Chromosomal location

Comments

1 liver 25-34 7p existence of both phosphorylated and unphosphorylated forms

2 CNS 32-34 2q preference for IGF-II

3 global 29-54 7p M.W. dependant on degree of glycosylation; b

4 bone/CNS/prostate 24 & 29 17q M.W. depends on glycosylationb

5 kidney 23 2q A factor in lung and bone development

6 ovary/prostate 30-32 12q Higher affinity for IGF-II than IGF-I

7 several sites 26 4q Also known as Mac25 or IGFBPrP1

a Primary site of production b M.W. (molecular weight)

79

The IGFBPs have a greater affinity for IGF than the IGF receptor itself. Thus IGF signalling is

controlled via competitive inhibition. The activity of IGFBPs is controlled by proteases such as

MMP-7 and MMP-9, however it appears that the nature of deactivating protease is tissue

dependant (157).

Clusterin is a highly abundant protein in serum (71) and is ubiquitous in tissue

distribution. Some biological processes involving clusterin include sperm maturation, tissue

differentiation, tissue remodelling, cell proliferation, and cell death. Clusterin is also involved

in several pathological states including neurodegenerative diseases and cancer (138, 149). In

fact, the protein is disregulated in many cancers including pancreatic, prostate, breast, and

esophageal cancer (19, 134, 175, 180). Interestingly, the expression of clusterin appears to be

cancer-dependant. In ovarian carcinoma, clusterin overexpression provides protection for

tumors against chemotherapeutic agents. This has been discussed in the preceding chapter (see

chapter 3, Discussion).

Adhesion molecules play a pivotal role in keeping the monolayer of surface epithelium

attached to the basement membrane in the ovary. Basement membranes are composed of many

components, including collagenous and non-collagenous proteins, proteoglycans, and growth

factors. Laminin is the major non-collagenous protein of the basement membranes. The

integrins are heterodimeric proteins composed of an α and β subunit that bind to laminin (74,

151). A positive correlation between integrin β4 and laminin expression in serous ovarian

carcinoma has been shown by Skubitz et al. (151). However no studies have been done looking

at integrin β4 expression in other forms of EOC. Therefore we looked at integrin β4 expression

in mucinous, clear-cell, and endometrioid ovarian carcinomas as well as serous carcinomas.

80

Intracellular adhesion molecule-5 (ICAM5), also known as telencephalin, is a member of

the ICAM-family of adhesion proteins. These proteins bind to leukocyte β2-integrins

(CD11/CD18). ICAM-5 is highly expressed in the mammalian forebrain, appears at the time of

birth, and is located in the soma and dendrites of neurons (54). ICAM5 was found in the cell

culture supernatants of ovarian cancer cell lines (63) and therefore is an interesting molecule to

explore in ovarian cancer for several reasons:

1. The molecule seems to be exclusive to neuronal tissue. A caveat though is that this

was confirmed in mouse models.

2. It can suppress T-cell activation

3. Ovarian cancer cell lines express ICAM5

4. Adhesion molecules play a pivotal role in cancer

Endothelial Protein C Receptor (EPCR) is a critical protein that regulates the

anticoagulant functions of activated protein C (aPC), which is a liver-derived serine protease

(52). Studies in mice suggest that EPCR is important for embryonic development (28, 62).

EPCR also regulates some of the anti-inflammatory and anti-apoptotic functions of activated

protein C (aPC). For example, EPCR may protect the lungs from severe inflammatory lung

diseases by mediating aPC anti-inflammatory activity. EPCR is normally found on the

endothelium of large blood vessels, however studies have shown expression in dendritic cells,

where it plays a role in innate immunity, and in neutrophils, where expression prevents

neutrophil chemotaxis (47). With respect to ovarian cancer, very little is known. Suzuki et.al.

reported that aPC increases migration of ovarian cancer cells and through an EPCR mediated

81

pathway, increases invasion (159). However, very little has been said whether or not primary

ovarian tumors express EPCR. In the present study, we address this question.

82

4.2 Materials and Methods

4.2.1 Materials

Antibodies to ADAM15, EPCR, Clusterin, ICAM-5, IGFBP5, IGFBP7, and integrin β4 were

purchased from R&D Systems.

4.2.2 Tumor specimens

Human ovarian tissue samples were taken from formalin-fixed, paraffin embedded

samples. The tissues included normal ovarian surface epithelium, and twenty-one cases of

epithelial ovarian cancer (EOC): six serous, five mucinous, five clear cell and five endometrioid

adenocarcinomas. The carcinomas were graded according to the International Federation of

Gynecology and Obstetrics (FIGO) histological grade.

4.2.3 Immunostaining

Immunohistochemical staining was performed on 3µm thick paraffin embedded sections

of tissues fixed in buffered formalin. Staining of normal tissue, cancerous tissue, and positive

controls were performed concurrently. Staining procedures included deparaffinization in warm

xylene for 5 min with two changes of xylene at room temperature, followed by rehydration by

transfer through graded alcohols and then rinsing with distilled water. The Trilogy antigen

retrieval system (Cell Marque) was used for one hour in order to expose the antigen epitopes.

After 20 minutes at room temperature and rinsing with distilled water, the sections were put in

3% H202 for 10 min in darkness. After washing with tap water, the sections were dipped twice

for 5 min in Tris buffer saline (TBS), incubated with the following antibodies, at the following

dilutions: mouse monoclonal ADAM15 (1:10), goat polyclonal Clusterin (1:1000), goat

polyclonal EPCR (1:400), mouse monoclonal ICAM-5 (1:10), monoclonal mouse IGFBP-5

83

(1:20), goat polyclonal IGFBP-7 (1:200), and mouse monoclonal integrin β4 (1:20) for 30 min,

rinsed with TBS for 10 min and then incubated with the Envision™ detection system

peroxidase/DAB+, Rabbit/Mouse (DAKO Cytomation, Denmark) for 30 min. Rinsing with

TBS for 10 minutes, incubation in diaminobenzidine (DAB) solution for 10 minutes at room

temperature and rinsing with tap water followed. The sections were then counterstained with

haematoxylin, dehydrated, cleared in xylene and mounted. Negative controls were performed

for all studied tissues by omitting the primary antibody

4.2.4 Evaluation of immunohistochemical staining

Staining was evaluated by an experienced Pathologist. Cytoplasmic/membranous

staining was evaluated as no staining (0), weak (1), moderate (2), or strong (3) based on

intensity. The percentage of each expression (weak, moderate, or strong) was assigned based on

the proportion of positive tumor cells to total tumor cells ranging from 0 to 100%. Percentages

for weak, moderate, and strong staining in every tumor section were multiplied by 1, 2 and 3

respectively. That is, if a tumor has no stain, the value is 0 (0 x 100). If it has strong staining in

100% of the section (the whole tumor is stained strongly) the value is 300 (3 x 100). If a tumor

section has 30% mild expression and 20% strong expression, then the value is 90 (0x50 + 1x30

+ 3x20). The obtained values (0-300) were categorized in four score groups with the following

ranges: 1=0-50, 2=51-150, 3=151-250, 4=251-300. A final score was obtained by using a two-

scale grouping, as follows: 1 (low expression)=0-150 and 2 (high expression)=151-300.

84

4.3 Results

The interpretation of staining intensity is highly subjective and may be affected by

storage time, variation in protocols, and fixation procedures. In addition, strength of antibody-

antigen interaction can also affect the intensity of staining. Therefore, we stained all tumour

tissue, normal tissue, and positive controls concurrently to keep the staining procedure constant

for each antibody. Staining intensity and scoring was performed by an experienced Pathologist.

4.3.1 ADAM15 expression

ADAM15 protein was detected in 5/6 serous carcinomas (Table 4.2). In the positive

cases, the expression was low regardless of the grade of the serous carcinoma. One tumor

section showed no staining for the protein. ADAM 15 was detected in all mucinous carcinomas

(5/5). Four specimens showed low expression and one showed high expression. The strength of

the staining did not correlate with tumor grade. All clear-cell carcinoma sections stained

positive for ADAM15. Again, the strength of the staining did not correlate with tumor grade.

However, 3/5 tumor sections (all FIGO stage II) showed high ADAM15 expression. Although

the expression was low, all five endometrioid carcinoma sections stained positive for ADAM15.

Tumor grade did not correlate with protein expression in these sections. Sections representative

of ADAM15 staining in the four subtypes of EOC are illustrated in Figure 4.1. Healthy ovarian

surface epithelium was also positive for ADAM15 protein, but expression was qualitatively

equivalent the cancer counterparts (Figure 4.2). Prostate cancer tissue was used as a positive

control for staining (result not shown). ADAM15 expression was observed in a predominately

cytoplasmic distribution pattern in all tissue sections including healthy surface epithelium.

85

4.3.2 Clusterin expression

Clusterin expression was observed in all serous carcinomas (Table 4.3). Four serous

carcinomas showed high expression and two samples showed low expression. No correlation

was observed between expression levels and tumor grade. All five mucinous carcinoma

specimens stained positive for clusterin. In all, three tumors showed high clusterin expression.

No correlation was observed between tumor grade and expression levels. Clusterin protein was

detected in all clear-cell carcinoma specimens. Two tumors showed high expression and three

showed low expression. Again, we did not see a correlation between tumor grade and levels of

expression. All endometrioid tumors stained positive for clusterin and all specimens showed

high clusterin expression. Representative sections of the four subtypes studied for clusterin

expression is illustrated in Figure 4.3. Lymph nodes were used as positive controls for staining

(data not shown). Clusterin expression showed a cytoplasmic and plasma membranous

distribution, as well as a granular distribution. Healthy ovarian surface epithelium showed

variable clusterin expression (Figure 4.4). No significant difference was found in clusterin

expression among the four histological types. Inflammatory cells, stromal cells, and the

endothelium within the ovary also showed clusterin expression.

86

Table 4.2: ADAM15 expression in ovarian tumors (proportion of positive cases)

Histological type ADAM15 expression positivea/total cases

Number of cancer cases showing staining

No/Low expressionb

High expression

Serous (n=6) 5/6 6 0

Endometrioid (n=5) 5/5 4 1

Clear cell (n=5) 5/5 2 3

Mucinous (n=5) 5/5 4 1

a Cases with at least weak staining were defined as ADAM15 positive b Includes tissues with no staining (ADAM15 negative)

87

Figure 4.1: Immunohistochemical expression of ADAM15 in the four major types of

epithelial ovarian cancer: A. Serous adenocarcinoma (x400), B. Mucinous adenocarcinoma

(x400), C. Clear cell adenocarcinoma (x100), D. Endometrioid adenocarcinoma (x200). Blue

staining is hematoxylin staining of nuclei. Brown staining shows the presence of ADAM15.

Arrows point to tumor cells expressing ADAM15.

88

Figure 4.2: Immunohistochemical expression of ADAM15 in normal surface epithelium

(400X). Arrow points to surface epithelial cells that are positive for ADAM15 expression

(brown staining). ‘Str’ designates stromal tissue. Blue colour is hematoxylin stain.

89

Table 4.3: Clusterin expression in ovarian tumors (proportion of positive cases)

Histological type Clusterin expression positivea/total cases

Number of cancer cases showing staining

No/low expressionb

High expression

Serous (n=6) 6/6 2 4

Endometrioid (n=5) 5/5 0 5

Clear cell (n=5) 5/5 3 2

Mucinous (n=5) 5/5 2 3

a Cases with at least weak staining were defined as clusterin positive b Includes tissues with no staining (clusterin negative)

90

Figure 4.3: Immunohistochemical expression of clusterin in the four major types of

epithelial ovarian cancer: A. Serous adenocarcinoma, B. Mucinous adenocarcinoma, C. Clear

cell adenocarcinoma, D. Endometrioid adenocarcinoma. All magnifications are x400, except B,

x200. Arrows point to tumor cells that are positive for clusterin (brown staining). Cell nuclei

were stained with hematoxylin (blue colour).

91

Figure 4.4: Immunohistochemical expression of clusterin in normal surface epithelium

(400X). Arrow depicts normal surface epithelial cells. Brown staining shows clusterin

expression. Blue stain (hematoxylin) shows cell nuclei. ‘Str’ designates stromal tissue.

92

4.3.3 EPCR Expression

EPCR protein was detected in all serous carcinomas (Table 4.4). In the positive cases,

three sections (50%) showed high expression. Expression levels of EPCR protein did not

correlate with tumor grade. EPCR was detected in all mucinous carcinomas (5/5). All

specimens showed low expression. The mucinous sections were either grade I or grade II

tumors. All but one clear-cell carcinoma section stained positive for EPCR protein. Although

expression levels were relatively low, the sections that stained positive were all from grade II

tumors. All but one endometrioid carcinoma sample stained positive for EPCR. Three sections

showed low expression, one showed high expression, and one showed no expression.

Expression levels did not correlate with tumor grade. Sections representative of EPCR staining

in the four subtypes are illustrated in Figure 4.5. Healthy ovarian surface epithelium was also

positive for EPCR protein, but the staining was qualitatively equivalent to or greater than the

cancer sections (Figure 4.6). We used liver tissue as a positive control for staining (result not

shown). EPCR expression was observed in both a granular cytoplasmic and membranous

distribution pattern in all tissue sections including healthy surface epithelium. As expected,

endothelial tissue was positive for EPCR.

93

Table 4.4: EPCR expression in ovarian tumors (proportion of positive cases)

Histological type EPCR expression positivea/total cases

Number of cancer cases showing staining

No/low expressionb

High expression

Serous (n=6) 6/6 3 3

Endometrioid (n=5) 4/5 4 1

Clear cell (n=5) 4/5 4 0

Mucinous (n=5) 5/5 5 0

a Cases with at least weak staining were defined as EPCR positive b Includes tissues with no staining (EPCR negative)

94

Figure 4.5: Immunohistochemical expression of EPCR in the four major types of epithelial

ovarian cancer: A. Serous adenocarcinoma (x400), B. Mucinous adenocarcinoma (x400), C.

Clear cell adenocarcinoma (x200), D. Endometrioid adenocarcinoma (x200). Arrows point to

tumor cells that are positive for EPCR (brown staining). Cell nuclei were stained with

hematoxylin (blue colour).

95

Figure 4.6 : Immunohistochemical expression of EPCR in normal surface epithelium

(200X). Arrow depicts normal surface epithelial cells. Brown staining shows EPCR

expression. Blue stain (hematoxylin) shows cell nuclei. ‘Str’ designates stromal tissue.

96

4.3.4 ICAM 5 Expression

ICAM5 protein was detected in all serous carcinomas (Table 4.5). Two grade II tumors

showed high expression of ICAM5, but we did not see a correlation between tumor stage and

expression levels. All mucinous carcinomas expressed ICAM5. Two tumor sections (one

grade I and one grade II) showed high expression. The strength of the staining did not correlate

with tumor grade. All clear-cell carcinoma sections stained positive for ICAM5, with one

section showing high expression. Again, the strength of the staining did not correlate with

tumor grade. Four endometrioid tumor sections showed low expression, and one (grade III)

showed high expression. Tumor grade did not correlate with protein expression in these

sections. Sections representative of ICAM5 staining in the four subtypes are illustrated in

Figure 4.7. Healthy ovarian surface epithelium showed negative to mild expression of ICAM5

and therefore was qualitatively less than cancer sections (Figure 4.8). We used breast cancer

tissue as a positive control for staining (result not shown). ICAM5 expression was observed in

a predominately cytoplasmic and membranous distribution pattern. Smooth muscle cells of the

ovary also expressed ICAM5.

97

Table 4.5: ICAM5 expression in ovarian tumors (proportion of positive cases)

Histological type ICAM5 expression positivea/total cases

Number of cancer cases showing staining

No/low expressionb

High expression

Serous (n=6) 6/6 4 2

Endometrioid (n=5) 5/5 4 1

Clear cell (n=5) 5/5 4 1

Mucinous (n=5) 5/5 3 2

a Cases with at least weak staining were defined to be ICAM5 positive b Includes tissues with no staining (ICAM5 negative)

98

Figure 4.7: Immunohistochemical expression of ICAM5 in the four major types of

epithelial ovarian cancer: A. Serous adenocarcinoma, B. Mucinous adenocarcinoma, C. Clear

cell adenocarcinoma, D. Endometrioid adenocarcinoma. All magnifications are x400. Arrows

point to tumor cells that are positive for ICAM5 (brown staining). Cell nuclei were stained with

hematoxylin (blue colour).

99

Figure 4.8: Immunohistochemical expression of ICAM5 in normal surface epithelium

(400X). Arrow depicts normal surface epithelial cells. Normal surface epithelial cells are not

positive for ICAM5, hence no brown stain. Blue stain (hematoxylin) shows cell nuclei. ‘Str’

designates stromal tissue.

100

4.3.5 IGFBP5 Expression

IGFBP5 protein was detected in three serous carcinomas (Table 4.6) but the expression

was low. Four mucinous carcinomas were positive for IGFBP5 protein with one having high

levels of expression. Two of the clear cell carcinoma specimens (tumor grade II) stained

positive for IGFBP5 and both showed high expression. Three of the endometrioid cancer

specimens stained positive for IGFBP5 with one showing high expression. Mild to moderate

expression was observed in normal surface epithelium. Representative samples from each EOC

subtype and normal surface epithelium showing IGFBP5 expression are illustrated in Figures

4.9 and 4.10. Placental tissue was used as a positive control (data not shown). Expression

followed a cytoplasmic and membranous distribution pattern. A correlation between tumor

stage and expression levels could not be established, as sample size was low. Reliable

statistical results could not be established to compare expression differences between the various

EOC subtypes as the number of samples were low. However, from a qualitative perspective,

serous carcinomas showed the weakest expression among the subtypes. Healthy ovarian

stromal tissue was also positive for IGFBP5 expression (data not shown).

4.3.6 IGFBP7 Expression

IGFBP7 protein was detected in all serous carcinomas (Table 4.7). All but one tumor

showed low expression. The sole high expressing tumor was a grade III tumor. A correlation

between tumor grade and expression levels could not be established. All mucinous carcinomas

were positive for IGFBP7 with low expression levels. All clear-cell carcinoma sections stained

positive for IGFBP7, with 3/5 sections showing high expression.

101

Table 4.6: IGFBP5 expression in ovarian tumors (proportion of positive cases)

Histological type IGFBP5 expression positivea/total cases

Number of cancer cases showing staining

No/low expressionb

High expression

Serous (n=6) 3/6 6 0

Endometrioid (n=5) 3/5 4 1

Clear cell (n=5) 2/5 3 2

Mucinous (n=5) 4/5 3 2

a Cases with at least weak staining were defined as IGFBP5 positive b Includes tissues with no staining (IGFBP5 negative)

102

Figure 4.9: Immunohistochemical expression of IGFBP5 in the four major types of

epithelial ovarian cancer: A. Serous adenocarcinoma, B. Mucinous adenocarcinoma, C. Clear

cell adenocarcinoma, D. Endometrioid adenocarcinoma. All magnifications are x400, except B,

which is x200. Arrows point to tumor cells that are positive for IGFBP5 (brown staining). Cell

nuclei were stained with hematoxylin (blue colour).

.

103

Figure 4.10 : Immunohistochemical expression of IGFBP5 in normal surface epithelium

(200X). Arrow depicts normal surface epithelial cells. Brown staining shows IGFBP5

expression. Blue stain (hematoxylin) shows cell nuclei. ‘Str’ designates stromal tissue.

104

The strength of the staining did not correlate with tumor grade, but the only grade III clear cell

tumor in the set showed high expression. All five endometrioid tumor sections were positive

for IGFBP7, with one grade II tumor and one grade III tumor showing high expression.

Sections representative of IGFBP7 staining in the four EOC subtypes are illustrated in Figure

4.11. Healthy ovarian surface epithelium showed negative to mild expression of IGFBP7 and

therefore was qualitatively less than cancer sections (Figure 4.12). IGFBP7 expression showed

a cytoplasmic and membranous distribution. Other ovarian tissues that tested positive for

IGFBP7 were endothelial and stromal tissue. We used endothelial tissue as a positive control

for staining.

4.3.7 Integrin β4 Expression

Integrin β4 protein was detected in 5/6 serous carcinomas (Table 4.8) with one showing

high expression. We did not obtain reliable data in mucinous carcinomas for integrin β4 protein

though one section was positive for high expression. Two of the clear cell carcinoma specimens

stained positive for integrin β4 with both showing low expression. Four endometrioid cancer

specimens stained positive for integrin β4 with one showing high expression. Normal surface

epithelium showed variable expression. Representative samples from each EOC subtype and

normal surface epithelium showing integrin B4 expression is illustrated in Figures 4.13 and

4.14. A squamous cell carcinoma was used as a positive control (data not shown). Expression

followed a cytoplasmic and membranous distribution pattern. A correlation between tumor

stage and expression levels was not established, as sample size was low. Healthy ovarian

stromal tissue was also positive for integrin B4 expression (data not shown).

105

Table 4.7: IGFBP7 expression in ovarian tumors (proportion of positive cases)

Histological type IGFBP7 expression positivea/total cases

Number of cancer cases showing staining

No/low expressionb

High expression

Serous (n=6) 6/6 5 1

Endometrioid (n=5) 5/5 3 2

Clear cell (n=5) 5/5 2 3

Mucinous (n=5) 5/5 5 0

a Cases with at least weak staining were defined to be IGFBP7 positive b Includes tissues with no staining (IGFBP7 negative)

106

Figure 4.11: Immunohistochemical expression of IGFBP7 in the four major types of

epithelial ovarian cancer: A. Serous adenocarcinoma, B. Mucinous adenocarcinoma, C. Clear

cell adenocarcinoma, D. Endometrioid adenocarcinoma. All magnifications are x200, except A,

which is x400. Arrows point to tumor cells that are positive for IGFBP7 (brown staining). Cell

nuclei were stained with hematoxylin (blue colour).

107

Figures 4.12: Immunohistochemical expression of IGFBP7 in normal surface epithelium

(200X). Arrow depicts normal surface epithelial cells. Brown staining shows IGFBP57

expression. Blue stain (hematoxylin) shows cell nuclei. ‘Str’ designates stromal tissue.

108

Table 4.8: Integrin β4 expression in ovarian tumors (proportion of positive cases)

Histological type Integrin β4 expression positivea/total cases

Number of cancer cases showing staining

No/low expressionb

High expression

Serous (n=6) 5/6 5 1

Endometrioid (n=5) 4/5 5 0

Clear cell (n=5) 2/5 5 0

Mucinous (n=5) 1/5 5 1

a Cases with at least weak staining were defined to be integrin β4 positive b Includes tissues with no staining (Integrin β4 negative)

109

Figure 4.13: Immunohistochemical expression of Integrin β4 in the four major types of

epithelial ovarian cancer (arrows): A. Serous adenocarcinoma, B. Mucinous adenocarcinoma,

C. Clear cell adenocarcinoma, D. Endometrioid adenocarcinoma. All magnifications are x200,

except A, which is x400. Arrows point to tumor cells that are positive for Integrin β4 (brown

staining). Cell nuclei were stained with hematoxylin (blue colour).

110

Figures 4.14: Immunohistochemical expression of Integrin β4 in normal surface

epithelium (200X). Arrow depicts normal surface epithelial cells. Brown staining shows

Integrin β4 expression. Blue stain (hematoxylin) shows cell nuclei. ‘Str’ designates stromal

tissue.

111

4.4 Discussion

Among gynaecological malignancies, EOC is the most lethal due to the advanced stage

at which the cancer is identified. The typical treatment option is bilateral salpingo-

oopherectomy and hysterectomy followed by taxane-based and platinum-based chemotherapy.

However, chemotherapy is confounded by the highly toxic side effects and some cases do not

respond to current chemotherapeutic strategies. Thus, identifying EOC specific proteins or

proteins that are overexpressed in EOC can lead to novel therapeutic strategies that are less toxic

and improve patient prognosis.

We studied seven potential EOC biomarkers, of which four (Clusterin, IGFBP5,

IGFBP6, and EPCR) were selected based on the criteria shown in the preceding chapter. That

is, these proteins were common to both the cell culture supernatants and ascites fluid from EOC

cases (93). The other three (ADAM15, Integrin β4, and ICAM5) were selected from the list of

extracellular and membrane proteins extracted from the discovery data set (see chapter 2).

These proteins were not found in the ascites study (93), presumably because the ascites study

focused on the sub-30 kDa proteome. ADAM15 is over 100 kDa (106), Integrin β4 is 150-250

kDa (87), and ICAM5 is over 100 kDa (163).

ADAM15 is a member of a disintegrin and metalloproteinase family (115). There are

two clearly defined biological roles for ADAM5, proteolysis and cell adhesion. As a protease,

one of the major proteins cleaved by ADAM15 is E-cadherin (118). Using breast cancer cell

lines, Najy et al. showed that the cleaved form of E-cadherin transactivated HER2/HER3, thus

resulting in increased migration and proliferation. The importance of E-cadherin in ovarian

tumors has been discussed by several investigators(103, 158). It remains to be seen if E-

112

cadherin is a substrate for ADAM15 in ovarian tumors. ADAM15 and its role in metastasis has

been documented before in prostate cancer. The loss of ADAM15 decreased metastasis to the

bone (117).

In the present study, we did not see a difference in ADAM15 expression between

healthy ovarian surface epithelial tissue and cancer tissue. However, an important point to note

is that the sample size (n=21 cancer tissues) was small in this study. Therefore whether or not

there was a statistically significant difference between healthy and cancerous tissue could not be

determined. Nevertheless, some of the cancerous tissue showed high expression of ADAM15,

especially in clear-cell tumors. Clear-cell tumors of the ovary are notoriously chemoresistant.

It remains to be seen whether or not ADAM15 is a suitable therapeutic target in ovarian cancer.

Another question that needs to be addressed is whether or not ADAM15 is more active in

ovarian cancer tissue compared to healthy tissue.

EPCR is a member of the protein C anticoagulant pathway (47). EPCR and its ligand,

activated protein C (aPC), have been implicated in tumor metastasis. However, these reports

have been contradictory. Bezuhly et al. demonstrated in murine melanoma metastasis models

that EPCR overexpressing mice were more resistant to metastatic disease compared to wild-type

mice(20). In addition, treatment of wild-type mice with recombinant aPC showed a 44%

reduction in lung metastases. In contrast, Beaulieu et al. showed that chemotaxis and invasion

increased with breast cancer cell lines as concentrations of aPC were increased (14). In

addition, blocking antibodies to EPCR, attenuated chemotaxis. It is unclear whether or not

EPCR plays a role in ovarian cancer metastasis. In the present study, EPCR was expressed in

both cancerous and normal tissue. However, we did not check for aPC status in cancer and

healthy ovarian tissue. It is possible that EPCR status remains the same in cancer and healthy

113

tissue, but the amount of aPC may be different. This needs to be addressed. To our knowledge,

this is the first report of EPCR expression in ovarian cancer tissue.

Clusterin’s role in ovarian cancer biology is unclear and it is only recently that its role

has been discussed in the literature. We found high clusterin expression in a majority of the

ovarian cancer cases, but expression did not correlate with tumor grade. Normal healthy tissue

expressed variable amounts of clusterin, and this is not a surprise as clusterin is a highly

abundant protein (71). The soluble (secreted) form of clusterin has been implicated in

chemoresistance and is prosurvival, whereas the nuclear form of this molecule is pro-apoptotic

(149). In our study, clusterin expression in ovarian cancer took a granular cytoplasmic

distribution. We hypothesize that some of this may be soluble clusterin contained in secretory

vesicles. Clusterin was also distributed around the plasma membrane. Again, this may be the

soluble form of clusterin.

The association between metastasis and integrin β4 has been revealed mostly by studies

in breast cancer models. Integrin β4 is associated with breast tumors originating from basal

cells, which demonstrate a more aggressive phenotype (132). Individuals that express both

integrin β4 and laminin-5 in their primary tumors have the poorest prognosis among breast

cancer patients (160). Tumors residing in lymph nodes that originated from integrin

β4−negative primary breast cancers are positive for integrin β4 (120). The connection between

integrin β4 and cancer metastasis is not limited to breast cancers. Integrin β4 expression is

found at the invasive front of gastric cancers (161) and is associated with lymph node metastasis

in papillary thyroid cancers (89). However, very little data has been shown for ovarian cancer.

One study showed that transformed epithelial cells isolated from ascites fluid from patients with

serous ovarian carcinoma had low expression of integrin β4 (151). The authors hypothesized

114

that this low expression may be a mechanism for tumor cell release. In this study, integrin β4

was found in all cancer subtypes, but with low expression. Normal tissue showed variable

expression. Since the sample sizes were small, a conclusion with supporting statistical analysis

could not be made regarding differences between cancerous and healthy tissue. To our

knowledge, this is the first report of integrin β4 reported in the other subtypes of ovarian cancer.

Given that integrin β4 is expressed and that it is an adhesion molecule that seems important to

cancer metastasis (as seen in other cancers), its role in the metastatic process of EOC needs to be

explored further.

The role of IGFBP5 in ovarian cancer is a relatively unexplored area. IGFBPs either

inhibit or enhance the actions of IGFs. Depending on the physiological/pathological status,

IGFBPs either stimulate or suppress cell proliferation. In the case of ovarian cancer, Wang et

al. showed that IGFBP5 expression was higher in high-grade serous carcinomas relative to low-

grade serous carcinomas, serous borderline tumors, benign cysts, and normal surface epithelium

(170). Similarly in our study, from a qualitative perspective, IGFBP5 expression was higher in

the cancer tissue compared to normal surface epithelium. Although we did not have a

sufficiently large sample size to obtain statistical power, in contrast to Wang et al., our study

showed that mucinous and clear-cell carcinomas showed high expression of IGFBP5. This may

be due to a difference in tumor grade between Wang’s mucinous and clear-cell carcinoma sets

and ours. In this study, grade I and grade II mucinous and clear-cell carcinoma samples were

used. It was unclear as to the grade/stage of the mucinous and clear-cell samples used by Wang

et al.

115

The role of IGFPB7 in ovarian cancer is also unclear. However, several studies have

hinted at a role for IGFBP7 in cancer. IGFBP7 binds preferentially to type IV collagen and

appears to be co-expressed in tumor-associated endothelium in colorectal carcinoma (155).

Some have hypothesized that IGFBP7 is a tumor suppressor (84, 92). In breast cancer cell lines,

Burger et al. reported the allelic loss of heterozygosity on chromosome 4q which is the location

of IGFBP7 (24). In the same study, invasive breast tumors were negative for IGFBP7 whereas

normal and benign samples showed strong IGFBP-7 expression using immunohistochemistry.

In contrast, in our study, normal surface epithelium showed no staining or very mild staining at

best, whereas all tumor samples (especially clear cell and endometrioid) showed higher

expression than the healthy tissue. It is possible that the biological function of IGFBP7 is

different in ovarian cancer compared to breast cancer.

ICAM5 in ovarian tumors was variable in that some tissues showed low expression and

others showed high expression. Overall, the cancer tissues expressed ICAM5 at higher levels

than healthy normal surface epithelium. A literature review combining ICAM5 and cancer as

the search terms showed only one study that indicated a role for ICAM5 in neoplasia (109). In

that study, ICAM-5 transcripts were detected in cell lines established from primary head and

neck, colon, thyroid, cervical, pancreatic, skin, and adenoid cystic carcinomas. Downregulation

of ICAM-5 using siRNA inhibited cell proliferation. In addition, primary squamous carcinomas

of oral mucosa showed higher expression of ICAM-5 relative to the matched normal healthy

tissue. Carcinomas with high ICAM-5 expression had higher incidences of perineural invasion.

Given that ICAM-5 may play a role in invasion and that it is indeed an adhesion molecule, it is

possible that ICAM-5 may contribute to the metastatic potential of ovarian carcinomas. This is

116

speculative and more experiments need to be conducted to explore a relationship between

ICAM5 and variables including tumor grade, invasiveness, and tumor subtype.

In summary, normal healthy surface epithelium showed low expression relative to the

cancer samples for all proteins except ADAM15 and EPCR. Although normal tissue showed

equivalent expression of ADAM15 and EPCR, we cannot rule out roles for these two proteins in

ovarian cancer. Given that ADAM15 is a part of a large superfamily of proteins that have been

implicated in other cancers, one needs to also explore the other ADAMs , such as ADAM17 and

ADAM 10, in the context of ovarian cancer. To our knowledge, this study is the first one

showing ADAM15, EPCR, IGFBP7, and ICAM5 protein expression in ovarian cancer.

Furthermore we also showed integrin β4 protein expression in clear-cell, mucinous, and

endometrioid carcinomas, which have not been documented in the literature. Given that these

proteins are indeed expressed in ovarian cancer and that the expression is qualitatively higher in

cancer compared to normal, correlative studies looking at protein expression, tumor grade, and

tumor subtype using a sufficiently large sample set to achieve statistical power are warranted.

117

Chapeter 5: Antibody Production and Immunoassay Development

for NPC2

118

5.1 Introduction

NPC2 (Niemann-Pick disease type C2 protein), also known as HE1 (epididymal

secretory protein E1), is a small soluble 151 amino-acid glycoprotein. It contains a 19 amino

acid signal peptide and was first characterized as a major secretory protein in the human

epididymis (88). The NPC2 gene has been mapped to chromosome 14q24.3 and is 13.5kb long.

The gene itself contains five exons that range from 78 to 342 bp in size (121).

The protein has been identified in secretory fluids such as milk, bile, and epididymal

fluid (88, 91, 121). NPC2 is more known for its role in Niemann-Pick Type C disease where a

mutation in the gene leads to an autosomal-recessive lipid storage disease (121). This storage

disease appears to be related to intracellular cholesterol homeostasis. Along with NPC1

(Niemann-Pick disease type C1 protein), NPC2 works to maintain cholesterol homeostasis (29,

110, 167, 179). NPC2 has an affinity for cholesterol in the micromolar range and binds with a

1:1 stoichiometry (176). The majority of cases of Niemman-Pick involve the mutation of the

NPC1 gene whereas 5% of the cases involve NPC2. In Niemann-Pick disease, an abnormal

accumulation of cholesterol is seen in cells, presumably due to problems with cholesterol

trafficking.

The primary structure of NPC2 contains three potential glycosylation sites at Asn-19,

Asn-39, and Asn-116. Asn-19 appears to remain unglycosylated whereas Asn-39 is linked to an

endo H-sensitive oligosaccharide, and Asn-116 may remain unglycosylated, but if glycosylated

then its either an endo-H resistant or sensitive oligosaccharide (101). The glycosylation pattern

is heterogeneous and appears tissue specific (152).

119

In using our integrated approach to biomarker discovery (see Chapter 3), NPC2 was one

of our final 51 candidates. It was expressed by the four cell lines studied and was also found in

ascites fluid (63). Therefore, we hypothesized that NPC2 may be candidate biomarker for

ovarian cancer. The aim of this study was to see if NPC2 is indeed a suitable biomarker.

120

5.2 Materials and Methods

5.2.1 Biological specimens

Semen was obtained from patients of the fertility clinic. The collection was performed with

informed consent and approval by the Institutional Review Boards of Mount Sinai Hospital and

the University Health Network. Samples were stored at -80oC until use.

5.2.2 NPC2 purification from seminal plasma

Approximately 100 ml of seminal plasma was centrifuged at 10,000 xg for 20 minutes.

The supernatant was collected and dialyzed against HEPES buffer (20 mM HEPES, 100 mM

NaCl, pH 7.0) overnight. The sample was dialyzed twice against HEPES buffer to ensure that

the sample is at a pH was 7.

Two Tricorn 10/100 columns (GE Healthcare, Baie d’Urfe, PQ, Canada) were packed

with either SP sepharose™ fast flow (GE Healthcare) or Q sepharose™ fast flow (GE

Healthcare). Column packing was performed according to the manufacturer’s instructions. The

two columns were arranged in tandem, with the Q sepharose™ column leading into the SP

sepharose™ column. The tandem column was then installed on an ÄKTA™ FPLC System (GE

Healthcare).

The tandem column was first washed with 5 column volumes of 20 mM HEPES buffer

(pH 7.0) containing 100 mM NaCl followed by 5 column volumes of 20 mM HEPES buffer (pH

7.0) containing 1M NaCl. The column was then re-equilibrated with 20 mM HEPES buffer (pH

7.0) containing 100 mM NaCl. The aforementioned procedure was conducted to “activate” the

column. The dialyzed seminal plasma sample was injected into the tandem column at flow rate

121

of 1.0 ml/min and the flow-through was collected. The column was washed with an additional 2

column volumes of 20 mM HEPES containing 100 mM NaCl. The wash was combined with

the flow-through. Bound proteins were eluted using 20 mM HEPES buffer (pH 7.0) containing

1M NaCl and kept for later analysis.

The flow-through was then dialyzed against 12.5 mM ammonium acetate (pH 4.7)

overnight with two buffer exchanges to ensure that the sample was equilibrated well. A

prepacked SP sepharose™ fast flow column with a volume of 5 ml (GE Healthcare) was first

washed with 5 column volumes of 12.5 mM ammonium acetate (pH 4.7), then washed with 5

column volumes of 1M ammonium acetate (pH 4.7), and finally equilibrated with 10 column

volumes of 12.5 mM ammonium acetate (pH 4.7). The dialyzed sample was injected into the

column at a flow rate of 1.0 ml/minute. The column was washed with 2 column volumes of

12.5 mM ammonium acetate (pH 4.7). Column washes were kept for later analysis.

Sample elution was performed on the benchtop, using step gradients of 50 mM, 100 mM

150 mM, 200 mM, 250 mM, 300 mM, 350 mM, 400 mM, and 1M ammonium acetate. Three to

five fractions of each gradient were collected (one column volume per fraction). Each fraction

was separated by SDS-PAGE to see the purity. The gel-region between 20-27 kDa was excised

and processed for mass spectrometric analysis to determine which fractions contained NPC2.

5.2.3 In-Gel Digestion

Each excised gel band was first cut into small cubes (1 mm3) and placed in 1.5 ml

microcentrifuge tubes containing 100 µl of 50 mM ammonium bicarbonate. The ammonium

bicarbonate was aspirated and 100 µl of neat acetonitrile was added and incubated for 10

minutes at room temperature. The gel pieces were then centrifuged using a benchtop

122

microcentrifuge and the liquid was aspirated. Approximately 50 µl of 10 mM DTT (in 50 mM

ammonium bicarbonate) was then added and incubated in a water bath for 30 minutes at 56 0C.

Gel-pieces were then centrifuged and the DTT solution was aspirated. The microcentrifuge

tubes were then chilled to room temperature and 100 µl of neat acetonitrile was added and

incubated for 10 minutes. The gel-pieces were then centrifuged and the acetonitrile was

aspirated. Approximately 50 µl of 50 mM iodoacetamide (in 100 mM ammonium bicarbonate)

was added to the gel pieces and incubated for 20 minutes at room temperature in the dark.

Iodoacetamide was removed and 100 µl of neat acetonitrile was added to shrink the gel pieces.

Once the gel pieces were shrunk, approximately 75 µl of trypsin solution (20 µg trypsin

dissolved in 1 ml of 50 mM ammonium bicarbonate) was added and incubated overnight at

370C.

5.2.4 Mass spectrometric analysis

Supernatants from digested gel pieces were loaded onto a ZipTip C18 pipette tip

(Millipore; catalogue number ZTC18S096) and eluted in 4µl of Buffer B (90% acetonitrile

(ACN), 0.1% formic acid, 10% water, 0.02% Trifluoroacetic acid (TFA)). The eluate was

mixed with 80 µl of Buffer A, and 40 µl were injected via an autosampler into an Agilent 1100

series HPLC. The peptides were first injected onto a 2-cm C18 trap column (inner diameter,

200 µm), and then eluted from the trap column into a resolving 5-cm analytical C18 column

(inner diameter, 75 µm) with an 8 µm tip (New Objective). The LC setup was coupled online to

a 2-D linear ion trap (LTQ, Thermo Inc.) mass spectrometer using a nano-ESI source in data-

dependent mode. Each fraction was run on a 35-minute gradient. The eluted peptides were

subjected to MS/MS. DTAs were created using the Mascot Daemon (version 2.22) and

extract_msn. We used the following parameters for DTA creation: minimum mass, 300 Da;

123

maximum mass, 4000 Da; automatic precursor charge selection; minimum peaks, 10 per

MS/MS scan for acquisition; and minimum scans per group, 1.

For the product ion monitoring (PIM) assay, tryptic peptides were separated on a 2 cm

C18 trap column (inner diameter 200 µm). The peptides were eluted from the trap column onto

a resolving 5 cm analytical C18 column (inner diameter 75 µm) with a 15 µm tip (New

Objective). The LC setup was coupled online to a 2-D linear ion trap (LTQ, Thermo Inc.) mass

spectrometer using a nanoelectrospray ionization source (nano-ESI). Buffer A contained 0.1%

formic acid, 5% ACN, and 0.02% TFA in an aqueous water solution, and buffer B contained

90% ACN, 0.1% formic acid, and 0.02% TFA in water. The eluted peptides were analyzed by

tandem mass spectrometry (MS/MS) for identification purposes and by PIM to identify NPC2

specifically in positive-ion mode. A linear gradient was used with an injection volume of 40 µL,

which was loaded onto the column via an Agilent 1100 Cap-LC series autosampler. A 25 min

method was developed with a 5 min gradient and used for all experiments.

5.2.5 PIM assay design

LC-MS/MS analysis of semi-purified NPC2 was performed and the major ions observed

were further analyzed by MS/MS using the LTQ. The peptides showing the greatest signal

intensity were noted. To assist in the determination of single reaction monitoring (SRM)

transitions, the Global Proteome Machine database (33) was used to select the peptides of NPC2

that are frequently identified. Furthermore, an extensive search of the literature for mass

spectrometric data on NPC2 was performed to further assist in identifying the proteotypic

peptides of NPC2. Using single reaction monitoring (SRM) transitions that were available from

LC-MS/MS proteomic survey data and extensive testing of all parent peptides of NPC2, PIM

assays were developed for the precursor-fragment ion transition of the:

124

1. Doubly charged intact proteotypic peptide with m/z 922 (EVNVSPCPTQPCQLSK) to m/z

732 (PCQLSK), m/z 1315 (PCPTQPCQLSK), and m/z 1402 (SPCPTQPCQLSK).

2. Triply charged intact proteotypic peptide wit m/z 811

(AVVHGILMGVPVPFPIPEPDGCK) to the doubly charged m/z 628 (PFPIPEPDGCK),

doubly charged m/z 726 (PVPFPIPEPDGCK), and m/z (AVVHGILMGVPV)

A scan range of 250-2000 was used with a full scan type. A collision energy of 21% was used

for MS/MS. The results were searched with Mascot (Matrix Science) to confirm the identity of

the moiety.

5.2.6 Rabbit immunization

One female rabbit was immunized with 500 µg of NPC2 purified from seminal plasma.

Prior to immunization, a sample of pre-immune blood was taken for controls. A boost was

given 4 weeks after initial inoculation. The second boost was given 4 weeks after that and the

first titre check was performed within one week. Rabbit serum was confirmed for anti-NPC2

antibodies by direct ELISA and then was exsanguinated .

5.2.7 Antibody purification

Antibodies were purified from rabbit sera by protein A affinity purification. Protein A

purification was performed using the kit system, MAPS (Bio-Rad Laboratories, Hercules, CA,

USA), according to the manufacturer’s instructions. Antibody concentration was determined to

be 0.6 mg/ml by absorbance at 280 nm.

125

5.2.8 Western Blotting

Western blots for NPC2 were performed using our in-house purified rabbit polyclonal

antibody (Gunawardana Ab). A second rabbit anti-NPC2 polyclonal antibody (Lobel Ab) was

also used in western blots (a gift from Dr. Peter Lobel). Membranes were blocked for 1 hr at

room temperature using 5% BSA in PBS-tween (0.1% v/v). Antibodies were diluted 1:1000 in

PBS containing 0.1% Tween and the membranes were incubated overnight at 4oC. Membranes

were washed using PBS-Tween (0.1% v/v). Goat anti-rabbit antibodies coupled to alkaline

phosphatase (1:10000 dilution in PBS containing 0.1% Tween and 5% BSA) was used as the

secondary antibody. Signals were developed and captured on X-ray film (GE Healthcare Life

Sciences) by using a chemiluminescent substrate (Diagnostics Product Corporation, Los

Angeles, CA, USA).

5.2.9 Biotinylation of detection antibody

Approximately 50 ng of polyclonal anti-NPC2 antibody was incubated with 50 ng of biotin in

0.5 M NaHC03 for 1 h. This was used as the detection antibody for the NPC2 assay.

5.2.10 Immunoassays

Sandwich-type ELISAs were constructed in-house using the polyclonal antibody for

antigen capture and a biotinylated polyclonal antibody for detection. White polystyrene

microtitre plates were coated with either 500ng/100µl polyclonal antibody in coating buffer

(50mM Tris buffer, 0.05% sodium azide, pH 7.8) and stored at room temperature overnight.

Fifty microlitres of protein calibrators or samples, and 50µl of assay buffer [50mM Tris, 6%

BSA, 0.01% goat IgG, 0.1% bovine IgG (Sigma-Aldrich Inc, St. Louis MO), 0.005% mouse

IgG (Fortron Bio Science Inc, Morrisville, NC), 0.05% sodium azide, pH 7.8] with 0.5M KCl

126

were added to wells and incubated for 90 min with shaking at room temperature. The plates

were washed 6 times with washing buffer (5mM Tris, 150mM NaCl, 0.05% Tween-20, pH 7.8).

Approximately 100µl of biotinylated detection antibody (125ng/ml in assay buffer containing

0.5M KCl) were added to each well and incubated for 1 h at room temperature with shaking.

The plates were then washed six times with the washing buffer. Approximately 100µl

(5ng/well) of alkaline phosphatase-conjugated (ALP) streptavidin (Jackson ImmunoResearch) in

sample buffer (6% BSA, 50 mM Tris, 0.06% sodium azide, pH 7.8) was added to each well and

incubated for 15 min with shaking at room temperature. The plates were washed 6 times with

the wash buffer, and then 100 µL of substrate buffer [0.1 mol/L Tris buffer, pH 9.1, containing

0.5 mmol/L diflunisal phosphate (DFP), 0.1 mol/L NaCl, and 1 mmol/L MgCl2] were added to

each well and incubated for 10 min with shaking at room temperature. Approximately 100µl of

developing solution (1 mol/L Tris base, 0.15 mol/L NaOH, 2 mmol/L TbCl3, 3 mmol/L EDTA)

were added to each well and incubated for 1 min with shaking at room temperature. The

fluorescence was measured with an EnVision™ 2103 time- resolved fluorometer (Perkin

Elmer).

5.2.11 Immunohistochemistry

Immunohistochemical staining was performed on 3µm thick paraffin sections of tissues

fixed in buffered formalin. Staining procedures included deparaffinization in warm xylene for 5

min with two changes of xylene at room temperature, followed by rehydration by transfer

through graded alcohols and then rinsing with distilled water. The Trilogy™ antigen retrieval

system (Cell Marque) was used for one hour in order to expose the antigen epitopes. After 20

min in room temperature and rinsing with distilled water, the sections were put in 3% H2O2 for

10 min in darkness. After washing with tap water, the sections were dipped twice for 5 min in

127

Tris-buffered saline (TBS), incubated with the in-house rabbit polyclonal NPC2 (1:500) for 30

min, rinsed with TBS for 10 min and then incubated with the Envision detection system

peroxidase/DAB+, Rabbit/Mouse (DAKO Cytomation, Denmark) for 30 min. Rinsing with TBS

for 10 min, incubation in diaminobenzidine (DAB) solution for 10 min in room temperature and

rinsing with tap water followed. The sections were then counterstained with haematoxylin,

dehydrated, cleared in xylene and mounted. Negative controls were performed for all tissues by

omitting the primary antibody. Please see section 4.2.4 for evaluation protocol.

128

5.3 Results

5.3.1 Analysis of Cell Culture Supernatants

Mass spectrometric analysis of cell culture supernatants showed that NPC2 was

indentified in the supernatants of all four cell lines described in chapter 3. Table 5.1

summarizes these results. At least 3 unique peptides of NPC2 were found in each replicate in

the HTB-75 cell line; two unique peptides in each replicate of TOV-112D, at least one unique

peptide in the replicates of TOV-21G; and at least 2 unique peptides in each replicate of the

RMUG-S cell line.

5.3.2 Analyzing complex fluids for NPC2

Given that ascites fluid bathes ovarian tumors, we hypothesized that NPC2 would be

found in the fluid. NPC2 was indeed identified by Kuk et al. in ascites fluid using multiple

fractionation steps followed by mass spectrometry (93). We obtained the same sample in which

NPC2 was identified by Kuk and separated 100 µg of protein by SDS-PAGE. Gel bands were

sliced and processed for mass spectrometry analysis. Our initial analysis did not detect NPC2

protein. We hypothesized that the complexity of ascites fluid, even with the crude purification

step of SDS-PAGE, was too great and that this complexity prevented the mass spectrometer

from detecting NPC2.

5.3.2 Development of Product Ion Monitoring Assay for NPC2.

Our previous results indicated that neither the LTQ nor the Orbitrap mass spectrometers

were able to detect NPC2 in a complex fluid using the general identification scan mode.

129

Table 5.1: Number of peptides of NPC2 identified in ovarian cancer cell lines.

Cell Line Replicate No. Number of Unique Peptides

1 4

2 5

HTB-75

3 3

1 2

2 2

TOV-112D

3 2

1 2

2 0

TOV-21G

3 4

1 2

2 3

RMUG-S

3 3

130

To increase the sensitivity of detection we then used the PIM assay as described by Kulasingam

et al. (97). However, Kulasingam’s PIM assay used a purified antibody to immunocapture and

enrich for the particular protein of interest prior to detection by the mass spectrometry. In our

case, we did not have a purified antibody for immunocapture. Thus, we conducted the PIM

assay without this antibody enrichment step. The following steps were used to develop the PIM

assay for NPC2:

1. In the PIM assay, an ion-trap mass spectrometer is set up so that a single tryptic

peptide (parent ion) corresponding to the protein of interest is selected and

fragmented. PIM on an ion-trap mass spectrometer is very similar to single reaction

monitoring (SRM) or multiple reaction monitoring (MRM) performed on triple-quad

instruments (3, 90). The resulting product ions are used to identify the protein.

These peptides are selected carefully such that:

i. The peptide (parent ion) is specific to the protein of interest and only

to that protein. Long peptides are preferable as they are more likely to

be specific to a protein of interest.

ii. The peptide is ionisable and can be detected clearly in the initial MS

scan prior to sequencing.

iii. A clear fragmentation spectrum can be obtained by MS/MS

2. NPC2-specific peptides identified in ascites fluid (60, 93), seminal plasma

(unpublished work), and cell culture supernatants were examined (Table 5.2) to

select the peptides that are repeatedly identified (proteotypic peptides (104)).

131

3. Two peptides (peptide A, m/z = 922 and peptide B, m/z = 811) were selected as the

proteotypic peptides on the ion-trap mass spectrometer (Figures 5.1 and 5.2).

4. A standard of semi-purified NPC2 was used to calibrate the mass spectrometer. The

standard was prepared by:

i. Separating 100 µg of total protein from seminal plasma using SDS-

PAGE.

ii. Since endogenous NPC2 is a protein with a mass that ranges from 20

kDa – 27 kDa, this region was excised, reduced, alkylated, and then

trypsin-digested prior to mass spectrometry analysis.

iii. The presence of NPC2 was confirmed by mass spectrometry

iv. We verified that peptides A and B (see step 3) were identified in our

standard.

132

Table 5.2: Tryptic peptides of NPC2 identified by mass spectrometry

Source

Cell Culture Supernatantsa

Ascites Fluidb Seminal Plasmac

AVVHGILMGVPVPFPIPEPDGCK

NQSLFCWEIPVQIVSHL

DCGSVDGVIK

EVNVSPCPTQPCQLSK

LVVEWQLQDDK

LVVEWQLQDDKNQSLFCWEIPV

QIVSHL

AVVHGILMGVPVPFPIPEPDGCK

EVNVSPCPTQPCQLSK

DCGSVDGVIK

LVVEWQLQDDKNQSLFCWEIPV

QIVSHL

AVVHGILMGVPVPFPIPEP

DGCK

EVNVSPCPTQPCQLSK

DCGSVDGVIK

LVVEWQLQDDK

a Peptides identified in cell culture supernatants b Peptides identified in Ascites fluid c Peptides identified in Seminal plasma

133

Figure 5.1: LC-MS/MS analysis of semi-purified NPC2 protein showing the identification

of proteotypic peptide A (upper panel) and the MS/MS daughter ions produced after

fragmentation of peptide A (lower panel). The arrows in the lower panel point to the three

daughter ions (y6, y11, and y12) used to identify NPC2 using the PIM assay.

134

Figure 5.2: LC-MS/MS analysis of semi-purified NPC2 protein showing the identification

of proteotypic peptide B (upper panel) and the MS/MS daughter ions produced after

fragmentation of peptide B (lower panel). The arrows in the lower panel point to the three

daughter ions (y112+, y132+, and b12) used to identify NPC2 using the PIM assay.

135

5.3.3 Optimizing the PIM assay for complex biological fluids

Seminal plasma was first fractionated using gel filtration chromatography.

Approximately 500 µl of seminal plasma was separated by gel filtration using a 60-minute

fractionation run. Fractions were collected from the 21-minute mark till the end of the run at.

To detect NPC2 in the fractions, we constructed a direct ELISA by coating each fraction on a

96-well plate. A polyclonal rabbit anti-NPC2 (a gift from Dr. Peter Lobel) was used as the

primary antibody, and a goat anti-rabbit monoclonal conjugated to alkaline phosphatase as the

secondary. Two peaks were observed, suggesting that there are two major species of NPC2 in

seminal plasma (Figure 5.3). The first peak elutes between fractions 32-36 and the second peak

elutes between fractions 39-43. These fractions were pooled, lyophilized, and resuspended in

PBS buffer. An aliquot of the sample was then separated by SDS-PAGE and gel bands were

excised from the 20-27 kDa region and processed for mass spectrometry. NPC2 was detected

by mass spectrometry using a general scan. It was not necessary to apply the PIM assay for this

sample as the levels of NPC2 in seminal plasma are relatively high.

We then separated 500 µl of ascites fluid by gel filtration chromatography. Fractions

32-36 and fractions 39-43 were pooled, lyophilized and resuspended in PBS buffer for SDS-

PAGE separation. Gel bands were cut from the 20-27 kDa region and processed for mass

spectrometry. We then used the PIM assay to detect NPC2. The NPC2 PIM standard was first

run to verify the elution times and fragmentation pattern for NPC2 peptides A and B (Figure

5.4). The elution time of peptide A was approximately 11.65 minutes and that of peptide B was

approximately 18.47 minutes. The daughter ions of peptide A (y6, y11, and y12) and peptides B

( y112+, y132+, and b12) were verified.

136

Figure 5.3: Direct ELISA for NPC2 in gel filtration fractions. Seminal plasma was

fractionated by gel filtration. The separation time was 60 minutes in length at 0.5 ml per

minute. Fractions were collected from the 21-minute mark to the end of the run. Each fraction

was coated directly on a 96-well ELISA plate and probed with a rabbit polyclonal anti-NPC2

antibody (Lobel Ab). Fluorescence (arbitrary units) at 615 nm is shown on the y-axis.

137

Figure 5.4: PIM assay using the NPC2 standard from seminal plasma for calibrating the

mass spectrometer. The arrow on the upper panel shows the elution time of peptide A (922

ion) and the arrows in the lower panel point to the daughter ions (y6, y11, y12) produced from

peptide A. The MS data was verified for NPC2 using Mascot and Scaffold. The lower half of

the upper panel (in red) shows the elution time for peptide B (811 ion).

138

The gel bands from the first ascites were then analyzed by PIM. Indeed, peptide A (elution time

of 11.2 minutes) and the corresponding y6, y11, and y12 daughter ions were detected, thus

confirming the presence of NPC2 in this sample of ascites fluid (Figure 5.5). We did not detect

peptide B at the elution time corresponding to the standard. There was a peak for a peptide with

an 811 m/z ratio eluting at approximately 15.2 minutes. However, the y112+, y132+, and b12

daughter ions were not seen for this peptide, thus suggesting that this was not peptide B of

NPC2.

PIM assay was repeated on four additional malignant ascites samples from ovarian

cancer patients. Samples were processed as mentioned above. Peptide A and the daughter

fragments were detected in all four ascites samples confirming the presence of NPC2 (Figure

5.5-5.9). Note, the elution time for peptide A is 11.6 minutes (± 0.4 seconds), which is

acceptable. Searching raw spectra against the MASCOT database followed by analysis of the

Mascot report using Scaffold verified the presence of NPC2 in all four ascites. Interestingly,

peptide B was not detected by the PIM assay. We suspect several reasons for this (see

Discussion).

5.3.4 Development of a high-throughput screening assay for NPC2.

Although our method was successful in detecting NPC2 in a complex biological fluid, the

procedure is not suitable as a high-throughput method for screening for the following reasons:

1. Large volume of sample is needed (500 µl). This becomes problematic when the

volume of valuable patient sera is low.

2. Each sample must be processed separately. This adds variability from sample to

sample. Given that there are several steps (chromatography, SDS-PAGE,

lyophilisation, etc.) with this procedure, sample handling becomes a crucial factor in

139

consistency.

140

Figure 5.5: PIM assay on malignant ovarian ascites sample 1. The upper panel shows a

peptide eluting at the 11.2 minute mark, but is within 0.4 seconds of the elution time of peptide

A seen for the NPC2 standard. The lower panel shows the fragmentation pattern of the 922-ion.

The y6, y11, and y12 ions are found (arrows). Note the lower half of the upper panel (in red)

shows the elution time for the 811-ion and this does not match that of peptide B for the NPC2

standard.

141

Figure 5.6: PIM assay on malignant ovarian ascites sample 2. The upper panel shows a

peptide eluting at the 11.98-minute mark, but is within 0.4 seconds of the elution time of peptide

A seen for the NPC2 standard. The lower panel shows the fragmentation pattern of the 922-ion.

The y6, y11, and y12 ions are found (arrows). The elution time for the 811-ion shown in the

lower part of the upper panel (in red) does not match that of peptide B in the NPC2 standard.

142

Figure 5.7: PIM assay on malignant ovarian ascites sample 3. The upper panel shows a

peptide eluting at the 11.78-minute mark, but is within 0.4 seconds of the elution time of peptide

A seen for the NPC2 standard. The lower panel shows the fragmentation pattern of the 922-ion.

The y6, y11, and y12 ions are found (arrows). The elution time for the 811-ion shown in the

lower part of the upper panel (in red) does not match that of peptide B in the NPC2 standard.

143

Figure 5.8: PIM assay on malignant ovarian ascites sample 4. The upper panel shows a

peptide eluting at the 11.72-minute mark, but is within 0.4 seconds of the elution time of peptide

A seen for the NPC2 standard. The lower panel shows the fragmentation pattern of the 922-ion.

The y6, y11, and y12 ions are found (arrows). The elution time for the 811-ion shown in the

lower part of the upper panel (in red) does not match that of peptide B in the NPC2 standard.

144

Figure 5.9: PIM assay on malignant ovarian ascites sample 5. The upper panel shows a

peptide eluting at the 11.61 minute mark, but is within 0.4 seconds of the elution time of peptide

A seen for the NPC2 standard. The lower panel shows the fragmentation pattern of the 922-ion.

The y6, y11, and y12 ions are found (arrows). The elution time for the 811-ion shown in the

lower part of the upper panel (in red) does not match that of peptide B in the NPC2 standard.

145

3. Procedure is cumbersome since only 5 samples can be processed within 2-3 days for

mass spectrometric analysis.

To screen at least 100 serum samples for NPC2, a more efficient high-throughput tool, such as a

sandwich-type ELISA, is needed. However, there were no commercially available antibodies

nor were there ELISA kits for NPC2. Therefore, we constructed an ELISA using a polyclonal

antibody produced in-house.

5.3.5 Development of Polyclonal anti-NPC2 antibody

To develop a working ELISA, a polyclonal antibody against NPC2 was produced.

Purified NPC2 from seminal plasma was used as the antigen. We used a modified version of

Lobel’s procedure for NPC2 purification (175). Approximately 100 ml of seminal plasma was

dialyzed against 20 mM HEPES buffer (pH 7.0) containing 100 mM NaCl overnight with two

buffer exchanges. The seminal plasma was first purified using a strong anion exchange and

strong cation exchange column in tandem. The columns were equilibrated to pH 7.0 with 20

mM HEPES buffer (pH 7.0) containing 100 mM NaCl. Equilibration of columns to a pH of 7.0

was necessary so that NPC2 would come out in the flow-through fraction whereas most of the

other proteins would bind to the ion-exchange resin. The flow-through fraction was then

dialyzed against 12.5 mM ammonium acetate buffer (pH 4.7) overnight with two buffer

exchanges. The dialyzed fraction was then fractionated using a strong cation exchange column

equilibrated to pH 4.7 using ammonium acetate. A 12.5 mM ammonium acetate buffer (pH 4.7)

was used as the binding buffer. Proteins bound to the resin were eluted in stepwise gradients

starting with 50 mM ammonium acetate. Three fractions were collected for each step gradient.

In all we used the following gradients: 50 mM, 100 mM, 150 mM, 200 mM, 250 mM, 300 mM,

146

350 mM, 400 mM, and 1 M ammonium acetate. Each fraction was separated by SDS-PAGE

(Figure 5.10). The gels were stained with Coomassie dye, and bands within the 18-28 kDa

range were excised and processed for verification. NPC2 eluted at step gradients of 100 mM

and 150 mM. For subsequent purifications of NPC2, we used step gradients of 100 mM and

150 mM ammonium acetate for purification (see Material and Methods for further details).

Fractions containing NPC2 were combined and concentrated to 1mg/ml of protein. The

purity of the sample was verified by both visualization on a SDS-PAGE gel and by mass

spectrometric analysis of the sample. The sample was then dialyzed against PBS (pH 7.2)

overnight with two buffer exchanges. Approximately 1 mg of protein was used to immunize one

female rabbit. The presence of anti-NPC2 antibodies in the first bleed was determined by direct

ELISA. Figure 5.11 shows the results of this experiment. A strong binding signal was seen

with all dilutions of the serum from the 1st bleed. The pre-immune serum consistently showed a

signal that was considerably lower than the signals obtained for the 1st bleed. Data are not

shown for the 1:1000 and 1:18000 dilution of the 1st bleed (our antibody). In some cases, the

signal of the pre-immune serum was equal to background (no antigen coated). The background

signal for the 1st bleed was higher than that of the pre-immune serum for the lower dilution of

serum. However, as one increased the dilution factor of the serum, the background signal of the

1st bleed dropped to the levels of the pre-immune serum, yet maintaining the relatively high

signals in the wells containing antigen. Several dilutions of seminal plasma were used as a

positive control.

147

Figure 5.10: Ion-exchange fractions separated by SDS-PAGE. Top panel: Lane 1,

molecular weight marker; lanes 2-6 correspond to fractions 1-5 collected by eluting with 100

mM ammonium acetate; lanes 7-11 correspond to fractions 1-5 collected by eluting with 150

mM ammonium acetate; lane 12 is a blank. Bottom panel: Lane 1, molecular weight marker;

lanes 2-6 correspond to fractions 1-5 collected by eluting with 200 mM ammonium acetate;

lanes 7-11 correspond to fractions 1-5 collected by eluting with 250 mM ammonium acetate;

lane 12 is a blank. Bands (arrows) were identified as NPC2 by MS analysis.

kDa

kDa

148

Figure 5.11: Determining the presence of anti-NPC2 antibodies in rabbit antisera.

Ninety-six well plates were coated with varying amounts of purified NPC2 (0, 10, 30, 100 ng).

Rabbit serum from the 1st bleed (our Ab) was used as the primary antibody at varying dilutions.

As a control, pre-immune sera and another anti-NPC2 antibody (Lobel Ab) was used. The

Lobel antibody was not used at 1:4000 dilution. Fluorescence (arbitrary units) at a wavelength

of 615 nm is shown on the y-axis.

149

Similar to the direct coating of purified NPC2, we detected a strong signal in all dilutions of

seminal plasma using the 1st bleed. The pre-immune serum did not detect NPC2 in the dilutions

of seminal plasma. We also used Lobel’s polyclonal antibody to verify the results of the direct

ELISA. Indeed, Lobel’s antibody detected the purified NPC2 protein.

We further verified the specificity of the new polyclonal by Western blotting (Figure

5.12). The western blot with the new polyclonal antibody showed two major bands in seminal

plasma as did the blot using the Lobel polyclonal. The new polyclonal did not detect purified

PSA (2 µg) but did detect 10 nanograms of purified NPC2, which we used as the immunogen

for the rabbit. The Lobel antibody also detected a single band in the purified sample of NPC2.

5.3.6 Construction of NPC2 immunoassay

Antibodies from rabbit anti-sera were affinity purified using protein A sepharose beads.

The purified polyclonal antibody was used as the coating antibody. An aliquot of the same

polyclonal antibody was biotinylated and then used as the detection antibody. The NPC2

immunoassay was optimized before testing serum samples. We used the following step-wise

approach to construct a sandwich-type immunoassay using a polyclonal-polyclonal

configurations.

1. NPC2 purified from seminal plasma was used as a standard, and assays were

optimized to produce a calibration curve that was close to being linear in the range of

the standards. Approximately 500 ng of anti-NPC2 antibody was used for coating.

Standards of 0, 0.5, 2, 5, 10, and 20 ng/ml NPC2 were used to construct the standard

curve (Figure 5.13)

150

Figure 5.12: Verifying the specificity of the new rabbit polyclonal anti-NPC2 antibody.

Panel A: Western blot for NPC2 using the new antibody developed in-house. Lane 1, seminal

plasma; lane 2, 2 µg of PSA, and lane 3, 10 ng of purified NPC2. Panel B: Western blot for

NPC2 using the Lobel Ab (a gift from Dr. Peter Lobel). Lane designations are identical to the

blot shown in panel A.

151

Figure 5.13: NPC2 calibration curve for the NPC2 sandwich-type ELISA. The new

polyclonal antibody was used as both the capture and detection antibody. Approximately 500

ng of antibody was used to capture antigen. The calibration standards were as follows: 0, 0.5,

2, 5, 10, and 20 ng/ml of NPC2. The y-axis shows fluorescence at a wavelength of 615 nm

(arbitrary units). R2 = 1.00. The experiment was performed three times and the graph is a

representation of one such experiment.

152

2. The ELISA was next tested for its efficacy in detecting endogenous protein. We

used ascites fluid that was positive for NPC2 as the test sample. Mass spectrometric

analysis verified the presence NPC2 in this ascites as described previously.

3. Lastly, we measured each analyte in serial dilutions of serum to examine the

relationship between the signal measured and the corresponding dilution.

4. We noted that the assay was very sensitive at the low end of the standard curve. That

is, the difference between background and the next highest standard (0.5 ng/ml

NPC2) was large.

5.3.5 Measuring NPC2 in serum

Serum levels of NPC2 were measured using NPC2 ELISA. Our initial measurement of

serum NPC2 indicated a high concentration of NPC2 in both normal serum and serum from

ovarian cancer patients. There was no difference between the two cases (data not shown). The

new polyclonal antibody is suitable for a sandwich type ELISA when using purified or a semi-

purified sample. However, the assay may not be suitable for a complex fluid such as serum

where the serum matrix may interfere with antibody binding. To address this issue, we lessened

the complexity of serum by fractionating samples by gel filtration prior to measuring NPC2

levels by ELISA. Serum from ovarian cancer patients (n=5) and normal healthy individuals

(n=5) were fractionated by gel-filtration chromatography. As positive controls, seminal plasma

and supernatants from CHO cells expressing recombinant NPC2 were also fractionated.

Fractions were collected starting at the 20-minute mark of the run to the 60-minute mark, at a

minute per fraction. NPC2 was measured in each fraction using the poly-poly sandwich-type

ELISA. Maximum amount of NPC2 was detected at approximately the 36-minute mark (Figure

5.14) in seminal plasma and then decreased towards the end of the run.

153

Figure 5.14: Levels of NPC2 in sera from patients with ovarian cancer and healthy

individuals. Five serum samples from ovarian cancer patients were fractionated by gel

filtration. Fractions were collected from the 20-minute mark untill the end of the run at the 60-

minute mark (middle graph). As controls, five serum samples from healthy individuals were

also fractionated (top graph). The x-axis shows fraction number and the y-axis shows the

concentration of NPC2 in each fraction. The bottom graph shows the fractionation runs of

seminal plasma and CHO supernatants containing secreted recombinant NPC2. For this graph,

the y-axis shows fluorescence at 615 nm (arbitrary units).

154

155

We suspect that the broadness of the peak is due to the large amounts of NPC2 present in

seminal plasma. Supernatants from CHO cells expressing recombinant NPC2 showed a similar

peak at the 36-minute mark but also showed a peak at the 40-minute mark. The NPC2 elution

profile of ovarian cancer sera and normal sera were very similar to each other and to that of the

CHO supernatants and seminal plasma. Indeed a peak at the 36-minute mark can be seen in

serum samples from both cancer and healthy cases. Although the peaks were greater in seminal

plasma and CHO supernatants, the general shape of the curves were same. The result of the

western blots (Figure 5.12) and the aforementioned ELISA indicates that our polyclonal

antibody is indeed detecting NPC2. There was no significant difference in NPC2 levels

between normal serum and ovarian cancer serum, although one normal serum sample showed a

relatively high amount of NPC2. However, the sample size is small and the ELISA needs to be

repeated with a larger set of samples.

Expression of NPC2 in ovarian cancer tissue by immunohistochemistry.

NPC2 protein was detected in all serous carcinomas by immunohistochemistry (Table

5.2), three of which showed high expression. Four mucinous carcinomas expressed low

amounts of NPC2 protein with one having no expression. All clear-cell carcinomas expressed

NPC2, with three showing high expression. All five endometrioid carcinoma specimens stained

positive for NPC2 with three showing high expression. No expression was seen in normal

surface epithelium. Representative samples from each EOC subtype and normal surface

epithelium showing NPC2 expression are illustrated in Figures 5.15 and 5.16. Epididymal

tissue was used as a positive control (data not shown). Expression followed a cytoplasmic

distribution pattern, but the staining appeared granular and dot-like. A correlation between

tumor stage and expression levels could not be established, as sample size was low.

156

Table 5.3: NPC2 expression in ovarian tumors (proportion of positive cases)

Histological type NPC2 expression positivea/total cases

Number of cancer cases showing staining

No/low expressionb

High expression

Serous (n=6) 6/6 3 3

Endometrioid (n=5) 5/5 3 2

Clear cell (n=5) 5/5 3 2

Mucinous (n=5) 4/5 5 0

a Cases with at least weak staining were defined as NPC2 positive b Includes tissues with no staining (NPC2 negative)

157

Figure 5.15: Immunohistochemical expression of NPC2 in the four major types of

epithelial ovarian cancer (arrows): A. Serous adenocarcinoma (x200), B. Mucinous

adenocarcinoma (x200), C. Clear cell adenocarcinoma (x400), D. Endometrioid

adenocarcinoma (x400). Arrows point to tumor cells that are positive for NPC2 (brown

staining). Cell nuclei were stained with hematoxylin (blue colour).

158

Figures 5.16: Immunohistochemical expression of NPC2 in normal surface epithelium

(400X). Arrow depicts normal surface epithelial cells. NPC2 expression in normal surface

epithelium is low (brown stain). Blue stain (hematoxylin) shows cell nuclei. ‘Str’ designates

stromal tissue.

159

Reliable statistical results could not be established to compare expression differences between

the various EOC subtypes, as the number of samples was low. However, given that the normal

surface epithelium stains negative, qualitatively there is a difference between cancer tissue and

controls. Other ovarian tissue that stained positive for NPC2 were stromal and endothelial

tissue, and inflammatory cells.

160

5.4 Discussion

NPC2 is a 20-28 kDa glycoprotein that is involved in cholesterol homeostasis (101,

121). The only report of NPC2 in relation to cancer was in thyroid papillary carcinoma, where

its expression was upregulated (6). In the same study, the investigators found that NPC2

appeared to be specific to papillary carcinoma and not other forms of thyroid cancer as its

expression was localized to the papillary projections. The authors concluded that NPC2 plays a

role in forming the papillary shape.

In our study, all four ovarian cancer cell lines expressed NPC2. More than one peptide

of this protein was detected by the mass spectrometer, suggesting that this protein is not

produced in extremely low amounts by malignant ovarian cells. Others have found NPC2 in

malignant ascites fluid taken from ovarian cancer patients (60, 93). We confirmed these

findings in five ascites fluid samples using the PIM assay. Peptide A with m/z 922

(EVNVSPCPTQPCQLSK) was the best proteotypic peptide among the two chosen in this study.

This particular peptide ionised well and was always detected when NPC2 was present regardless

of the complexity of the sample used. Peptide B (triply charged) with m/z 811

(AVVHGILMGVPVPFPIPEPDGCK) was not suitable, as it was not detected in ascites fluid

even though NPC2 was present. On further examination of MS results, we saw that peptide B

had a doubly charged state with an m/z of 1215 as well. Therefore, the peptide is divided

between the two charge states. It is possible that in these experiments, the concentration of

peptide B was divided between the m/z 811 and m/z 1215 variants. Consequently, the

concentration of peptide B with m/z 811 was too low for the mass spectrometer to detect and

hence not seen in the ascites samples. In spite of only peptide A working in the PIM assay, the

161

assay was successful in identifying NPC2 in malignant ascites fluid. These initial results

suggested that NPC2 was a good candidate for further study as a biomarker for ovarian cancer.

The PIM assay, though suitable for screening a few (5-10) samples, is cumbersome for a

screening study using a large set (>30) of samples. There were no commercial antibodies to

NPC2 available to construct a sensitive ELISA against NPC2. Therefore, we produced an in-

house rabbit polyclonal antibody against NPC2. The advantage of our technique was that we

used endogenous NPC2 protein purified from seminal plasma as the antigen, instead of a

recombinant protein. The polyclonal antibody was specific to NPC2 as seen in the western

blots. Using the in-house developed antibody, a poly-poly sandwich-type ELISA was

constructed. The assay showed linearity in the range of 0.5 ng/ml to 10 ng/ml of NPC2.

We measured NPC2 levels in sera from ovarian cancer and sera from healthy

individuals. We prefractionated the serum samples prior to testing the samples by ELISA. The

elution profile of NPC2 in the sera (cancer and normal) matched the elution profile of NPC2 in

seminal plasma and the recombinant NPC2 protein in CHO cell-supernatants. However, there

was no significant difference between healthy and cancer sera. Thus, the preliminary results

suggest that NPC2 is not a good serum-based biomarker to detect ovarian cancer. However, a

few points to remember are:

1. A larger sample size is necessary for a proper screening method. Therefore, a

conclusion with reliable statistical results cannot be made at this time.

2. The serum samples needed to be prefractionated before the ELISA. The purpose of

constructing an ELISA was to avoid the cumbersome prefractionation and sample

preparation steps involved with the PIM assay. This suggest that our antibody is not

162

robust for measuring NPC2 levels in serum, though it is very good for use in less

complex fluids such as cell culture supernatants, and for use in Western blots.

3. NPC2 is a protein with multiple isoforms (most likely due to glycosylation). The

elution profile shows two peaks and the Western blots also show two bands. At this

stage, one cannot determine if there is a particular species of NPC2 that is specific to

cancer.

Some of the aforementioned points can be addressed with the development of a better antibody,

preferably a monoclonal antibody. Such an antibody can be used to develop a new ELISA.

Glycoform specific antibodies can address the question of whether or not there is a species of

NPC2 that is specific to cancer.

As mentioned previously, some candidate proteins may be better markers for prognosis,

determining tumor grade/stage, or differentiating histological type. Whether or not NPC2 is

good for these cannot be determined yet. However, immunohistochemical analysis indicated

that NPC2 is expressed highly in EOC. We saw no expression in normal surface epithelium.

These results indicate that the presence of NPC2 is biologically relevant in EOC. These results,

however promising, are preliminary, and the role of NPC2 in EOC will be examined further.

163

Chapter 6: Summary and Future Directions

164

6.1 Summary

A proteomic platform using an integrated approach (combining multiple datasets) was

used to identify novel biomarkers to epithelial ovarian cancer. Overall, this thesis has provided

potential candidate proteins that may be used as either serological markers or

immunohistochemical markers to EOC. Below is a summary of the key findings of this study.

6.1.1 Key Findings

1. Chapter 2: Proteomic screen of cell culture supernatants

a. Two-dimensional liquid chromatography-tandem mass spectrometry (2D-LC-

MS/MS) approach was used to analyze the secretome of 4 ovarian cancer cell

lines. Each cell line represented a cancer of a particular histological type of

ovarian carcinoma (serous, mucinous, clear-cell, and endometrioid).

b. Over 2000 proteins were found in total, 420 proteins being extracellular or

plasma membrane proteins.

c. Known biomarkers of EOC including CA-125, HE4, KLK6, osteopontin, and

mesothelin were found using our strategy

2. Chapter 3: Selecting candidate proteins and verification of candidates (Figure 6.1)

a. A stringent set of criteria was used to select candidates. The list of 420

extracellular and plasma membrane proteins was narrowed down to a list of 51

proteins.

165

Figure 6.1: Flow chart representing the criteria used for candidate selection. See chapter 3,

section 3.3.1 for detailed explanation of criteria.

166

b. On the basis of antibody availability, nine proteins were selected from the list of

51 candidates. Immunoassays were constructed for 8 of the 9 candidates

i. Clusterin

ii. IGFBP4, -5, -6, and 7

iii. EPCR

iv. βIG-H3

v. Vasorin

c. Serum samples from patients with ovarian cancer and healthy individuals were

tested by ELISA for the 8 candidates. Clusterin and IGFBP6 showed a

difference in concentration between cancer and normals, with clusterin being

almost three times higher in cancer cases relative to healthy normals.

3. Chapter 4: The study of protein expression in EOC tissue by immunohistochemistry

a. Six proteins were studied for protein expression in ovarian cancer tissue:

i. Clusterin

ii. IGFBP5, and -7

iii. EPCR

iv. ICAM5

167

v. Integrin β4

b. A tissue set consisting of 6 serous, 5 mucinous, 5 clear-cell, and 5 endometrioid

ovarian carcinomas was used in combination with normal ovarian tissue

c. All proteins were expressed in ovarian carcinomas. Many of the candidates have

not been studied in ovarian cancer.

4. Chapter 5: NPC2

a. NPC2 was found in all four ovarian cancer cell lines and NPC2 was identified in

malignant ascites fluid from ovarian cancer patients.

b. A robust product ion monitoring (PIM) assay was developed to detect NPC2 in

complex fluids.

c. A rabbit polyclonal anti-NPC2 antibody was developed and an ELISA was

constructed

d. There was no difference in NPC2 levels in ovarian cancer sera compared to

normal sera.

e. NPC2 is expressed in high amounts in EOC tissue. No NPC2 expression was

seen in normal surface epithelium.

6.1.2 Proof of Hypothesis

1. Proteins secreted or shed by ovarian cancer cell lines are similar to those

secreted or shed by primary ovarian tumours.

168

We identified proteins such as CA-125, HE4, and KLK6 in the cell culture

supernatants. These proteins are also shed (CA-125) or secreted (e.g. HE4 and

KLK6) by primary ovarian tumours. Thus our approach shows that some proteins

that are shed/secreted by ovarian cancer cell lines are indeed similar to those that

are shed/secreted by primary tumours

2. These proteins can be identified by two-dimensional liquid chromatography-

coupled mass spectrometry (2D-LC MS).

We identified proteins such as CA-125, HE4, Mesothelin, and KLK6 by 2D-LC MS.

These proteins are well-studied biomarkers of ovarian cancer. Therefore the

approach presented in this study is a valid approach for biomarker discovery.

3. These proteins can be measured in biological fluids such as serum using

antibody based immunoassays and/or mass spectrometry-based single

reaction monitoring/multiple reaction monitoring assays.

Candidate proteins such as IGFBP4, -6, -7, Vasorin, Clusterin, and EPCR were first

identified in cell culture supernatants by 2D-LC MS and then identified in serum by

sandwich ELISA. Thus we showed that some proteins can be identified in serum

using sensitive immunoassays. We also showed that proteins (e.g. NPC2) can be

measured in biological fluids using product ion monitoring assays which are

identical to single reaction monitoring/multiple reaction monitoring assays.

4. Some proteins may serve as biomarkers for early detection or prognosis of

ovarian cancer.

169

Although we did not identify a good early detection marker, results from our

immunohistochemistry work and work shown in the literature suggests that certain

proteins such as NPC2 and clusterin may play a role in the pathological processes of

ovarian cancer and may therefore be markers for prognosis.

170

6.2 Future Directions

We have presented 51 proteins that may be potential biomarkers. In total, 9 proteins

were tested. NPC2 was a promising candidate. Although it seems to be a poor serological

marker, its expression in ovarian cancer tissue and absence in normal ovarian surface epithelium

suggests that NPC2 is biologically relevant to cancer. This aspect of NPC2 in relation to

ovarian cancer needs to be evaluated. Currently, with the exception of a study conducted in

papillary thyroid carcinoma, no studies have been done regarding NPC2 and cancer. This

protein should be tested in other cancers as well. We constructed a polyclonal-polyclonal

sandwich-type ELISA that needs further refinement. However the ideal ELISA would be one

that utilizes two highly specific monoclonal antibodies. These monoclonal antibodies need to be

developed for NPC2. A robust ELISA that is engineered for measuring serum levels of NPC2

will give a more solid answer to whether or not NPC2 is a good serological marker for ovarian

cancer.

As mentioned previously, 51 candidate proteins were found in this study. None of these

proteins has been evaluated in ovarian cancer. The reason for this is the lack of suitable

immunological reagents such as monoclonal and polyclonal antibodies and recombinant

proteins. Therefore, monoclonal and polyclonal antibodies need to be developed to measure the

serum levels of the remaining candidates in cancer and normal cases.

With respect to screening, some of the candidates studied in this thesis cannot be

dismissed yet. To draw a conclusion backed up with solid statistical power, one needs a large

sample set. Based on power analysis, a sample set containing at least 100 patient samples will

suffice. Proteins such as clusterin and NPC2 need to be evaluated using such a large sample set.

171

Furthermore, many of the candidates were indeed expressed in ovarian carcinoma. The

biological relevance of this expression needs to be studied in detail. Clues to their biological

significance can give hints as to which pathways or proteins can be used as therapeutic targets.

172

References

173

1. Akhmedkhanov A, Toniolo P, Zeleniuch-Jacquotte A, Kato I, Koenig K, and Shore R. Aspirin and epithelial ovarian cancer. Prev Med 33: 682-687, 2001. 2. Ammar H, and Closset J. Clusterin activates survival through the phosphatidylinositol 3-kinase/Akt pathway. J Biol Chem 283: 12851-12861, 2008.

3. Anderson L, and Hunter C. Quantitative mass spectrometric multiple reaction monitoring assays for major plasma proteins. Mol Cell Proteomics 5: 573-588, 2006.

4. Anderson N, and Anderson N. The human plasma proteome: history, character, and diagnostic prospects. Mol Cell Proteomics 1: 845-867, 2002.

5. Arai M, Utsunomiya J, and Miki Y. Familial breast and ovarian cancers. Int J Clin Oncol 9: 270-282, 2004.

6. Asakawa J, Kodaira M, Ishikawa N, Hirai Y, Nagataki S, Moatamed F, and Sugawara M. Two-dimensional complementary deoxyribonucleic acid electrophoresis revealing up-regulated human epididymal protein-1 and down-regulated CL-100 in thyroid papillary carcinoma. Endocrinology 143: 4422-4428, 2002.

7. Auersperg N, Edelson M, Mok S, Johnson S, and Hamilton T. The biology of ovarian cancer. Semin Oncol 25: 281-304, 1998.

8. Bast RJ. Status of tumor markers in ovarian cancer screening. J Clin Oncol 21: 200s-205s, 2003.

9. Bast RJ, Klug T, St John E, Jenison E, Niloff J, Lazarus H, Berkowitz R, Leavitt T, Griffiths C, Parker L, Zurawski VJ, and Knapp R. A radioimmunoassay using a monoclonal antibody to monitor the course of epithelial ovarian cancer. N Engl J Med 309: 883-887, 1983.

10. Bast RJ, and Knapp R. Use of the CA 125 antigen in diagnosis and monitoring of ovarian carcinoma. Eur J Obstet Gynecol Reprod Biol 19: 354-356, 1985.

11. Bast RJ, Siegal F, Runowicz C, Klug T, Zurawski VJ, Schonholz D, Cohen C, and Knapp R. Elevation of serum CA 125 prior to diagnosis of an epithelial ovarian carcinoma. Gynecol Oncol 22: 115-120, 1985.

12. Bast RJ, Urban N, Shridhar V, Smith D, Zhang Z, Skates S, Lu K, Liu J, Fishman D, and Mills G. Early detection of ovarian cancer: promise and reality. Cancer Treat Res 107: 61-97, 2002.

174

13. Bast RJ, Xu F, Yu Y, Barnhill S, Zhang Z, and Mills G. CA 125: the past and the future. Int J Biol Markers 13: 179-187, 1998.

14. Beaulieu LM, and Church FC. Activated protein C promotes breast cancer cell migration through interactions with EPCR and PAR-1. Exp Cell Res 313: 677-687, 2007.

15. Bell K, and Kurman R. A clinicopathologic analysis of atypical proliferative (borderline) tumors and well-differentiated endometrioid adenocarcinomas of the ovary. Am J Surg Pathol 24: 1465-1479, 2000.

16. Bell R, Petticrew M, Luengo S, and Sheldon T. Screening for ovarian cancer: a systematic review. Health Technol Assess 2: i-iv, 1-84, 1998.

17. Bengtsson S, Krogh M, Szigyarto C, Uhlen M, Schedvins K, Silfversward C, Linder S, Auer G, Alaiya A, and James P. Large-scale proteomics analysis of human ovarian cancer for biomarkers. J Proteome Res 6: 1440-1450, 2007.

18. Berglund L, Bjorling E, Oksvold P, Fagerberg L, Asplund A, Szigyarto CA, Persson A, Ottosson J, Wernerus H, Nilsson P, Lundberg E, Sivertsson A, Navani S, Wester K, Kampf C, Hober S, Ponten F, and Uhlen M. A genecentric Human Protein Atlas for expression profiles based on antibodies. Mol Cell Proteomics 7: 2019-2027, 2008.

19. Bettuzzi S, Davalli P, Astancolle S, Carani C, Madeo B, Tampieri A, and Corti A. Tumor progression is accompanied by significant changes in the levels of expression of polyamine metabolism regulatory genes and clusterin (sulfated glycoprotein 2) in human prostate cancer specimens. Cancer Res 60: 28-34, 2000.

20. Bezuhly M, Cullen R, Esmon C, Morris S, West K, Johnston B, and Liwski R. Role of activated protein C and its receptor in inhibition of tumor metastasis. Blood 113: 3371-3374, 2009.

21. Bhoola S, and Hoskins W. Diagnosis and management of epithelial ovarian cancer. Obstet Gynecol 107: 1399-1410, 2006.

22. Blobel C. ADAMs: key components in EGFR signalling and development. Nat Rev Mol Cell Biol 6: 32-43, 2005.

175

23. Borgono C, Kishi T, Scorilas A, Harbeck N, Dorn J, Schmalfeldt B, Schmitt M, and Diamandis E. Human kallikrein 8 protein is a favorable prognostic marker in ovarian cancer. Clin Cancer Res 12: 1487-1493, 2006.

24. Burger A, Zhang X, Li H, Ostrowski J, Beatty B, Venanzoni M, Papas T, and Seth A. Down-regulation of T1A12/mac25, a novel insulin-like growth factor binding protein related gene, is associated with disease progression in breast carcinomas. Oncogene 16: 2459-2467, 1998.

25. Campbell I, Russell S, Choong D, Montgomery K, Ciavarella M, Hooi C, Cristiano B, Pearson R, and Phillips W. Mutation of the PIK3CA gene in ovarian and breast cancer. Cancer Res 64: 7678-7681, 2004.

26. Cannistra S. Cancer of the ovary. N Engl J Med 351: 2519-2529, 2004.

27. Cao D, Ji H, and Ronnett B. Expression of mesothelin, fascin, and prostate stem cell antigen in primary ovarian mucinous tumors and their utility in differentiating primary ovarian mucinous tumors from metastatic pancreatic mucinous carcinomas in the ovary. Int J Gynecol Pathol 24: 67-72, 2005.

28. Castellino F, Liang Z, Volkir S, Haalboom E, Martin J, Sandoval-Cooper M, and Rosen E. Mice with a severe deficiency of the endothelial protein C receptor gene develop, survive, and reproduce normally, and do not present with enhanced arterial thrombosis after challenge. Thromb Haemost 88: 462-472, 2002.

29. Chang T, Chang C, Ohgami N, and Yamauchi Y. Cholesterol sensing, trafficking, and esterification. Annu Rev Cell Dev Biol 22: 129-157, 2006.

30. Chayka O, Corvetta D, Dews M, Caccamo A, Piotrowska I, Santilli G, Gibson S, Sebire N, Himoudi N, Hogarty M, Anderson J, Bettuzzi S, Thomas-Tikhonenko A, and Sala A. Clusterin, a haploinsufficient tumor suppressor gene in neuroblastomas. J Natl Cancer Inst 101: 663-677, 2009.

31. Chu C, and Rubin S. Screening for ovarian cancer in the general population. Best Pract Res Clin Obstet Gynaecol 20: 307-320, 2006.

32. Colombo N, Van Gorp T, Parma G, Amant F, Gatta G, Sessa C, and Vergote I. Ovarian cancer. Crit Rev Oncol Hematol 60: 159-179, 2006.

176

33. Craig R, Cortens JP, and Beavis RC. Open source system for analyzing, validating, and storing protein identification data. J Proteome Res 3: 1234-1242, 2004.

34. Cramer D, and Welch W. Determinants of ovarian cancer risk. II. Inferences regarding pathogenesis. J Natl Cancer Inst 71: 717-721, 1983.

35. Cravatt B, Simon G, and Yates Jr. The biological impact of mass-spectrometry-based proteomics. Nature 450: 991-1000, 2007.

36. Dauplat J, Hacker N, Nieberg R, Berek J, Rose T, and Sagae S. Distant metastases in epithelial ovarian carcinoma. Cancer 60: 1561-1566, 1987.

37. Diamandis E. Analysis of serum proteomic patterns for early cancer diagnosis: drawing attention to potential problems. J Natl Cancer Inst 96: 353-356, 2004.

38. Diamandis E. Mass spectrometry as a diagnostic and a cancer biomarker discovery tool: opportunities and potential limitations. Mol Cell Proteomics 3: 367-378, 2004.

39. Diamandis E, Yousef G, Soosaipillai A, and Bunting P. Human kallikrein 6 (zyme/protease M/neurosin): a new serum biomarker of ovarian carcinoma. Clin Biochem 33: 579-583, 2000.

40. Drapkin R, von Horsten H, Lin Y, Mok S, Crum C, Welch W, and Hecht J. Human epididymis protein 4 (HE4) is a secreted glycoprotein that is overexpressed by serous and endometrioid ovarian carcinomas. Cancer Res 65: 2162-2169, 2005.

41. Duffy M. Clinical uses of tumor markers: a critical review. Crit Rev Clin Lab Sci 38: 225-262, 2001.

42. Duffy M, McKiernan E, O'Donovan N, and McGowan P. Role of ADAMs in cancer formation and progression. Clin Cancer Res 15: 1140-1144, 2009.

43. Edwards D, Handsley M, and Pennington C. The ADAM metalloproteinases. Mol Aspects Med 29: 258-289, 2008.

44. Einhorn N, Bast RJ, Knapp R, Tjernberg B, and Zurawski VJ. Preoperative evaluation of serum CA 125 levels in patients with primary epithelial ovarian cancer. Obstet Gynecol 67: 414-416, 1986.

177

45. Einhorn N, Sjovall K, Knapp R, Hall P, Scully R, Bast RJ, and Zurawski VJ. Prospective evaluation of serum CA 125 levels for early detection of ovarian cancer. Obstet Gynecol 80: 14-18, 1992.

46. Enomoto T, Weghorst C, Inoue M, Tanizawa O, and Rice J. K-ras activation occurs frequently in mucinous adenocarcinomas and rarely in other common epithelial tumors of the human ovary. Am J Pathol 139: 777-785, 1991.

47. Esmon C. The endothelial protein C receptor. Curr Opin Hematol 13: 382-385, 2006.

48. Etzioni R, Urban N, Ramsey S, McIntosh M, Schwartz S, Reid B, Radich J, Anderson G, and Hartwell L. The case for early detection. Nat Rev Cancer 3: 243-252, 2003.

49. Faca V, Ventura A, Fitzgibbon M, Pereira-Faca S, Pitteri S, Green A, Ireton R, Zhang Q, Wang H, O'Briant K, Drescher C, Schummer M, McIntosh M, Knudsen B, and Hanash S. Proteomic analysis of ovarian cancer cells reveals dynamic processes of protein secretion and shedding of extra-cellular domains. PLoS One 3: e2425, 2008.

50. Fathalla M. Incessant ovulation--a factor in ovarian neoplasia? Lancet 2: 163, 1971.

51. Feeley K, and Wells M. Precursor lesions of ovarian epithelial malignancy. Histopathology 38: 87-95, 2001.

52. Foster D, and Davie E. Characterization of a cDNA coding for human protein C. Proc Natl Acad Sci U S A 81: 4766-4770, 1984.

53. Gagne J, Ethier C, Gagne P, Mercier G, Bonicalzi M, Mes-Masson A, Droit A, Winstall E, Isabelle M, and Poirier G. Comparative proteome analysis of human epithelial ovarian cancer. Proteome Sci 5: 16, 2007.

54. Gahmberg C, Tian L, Ning L, and Nyman-Huttunen H. ICAM-5--a novel two-facetted adhesion molecule in the mammalian brain. Immunol Lett 117: 131-135, 2008.

55. Geisler J, Buller E, and Manahan K. Estrogen receptor alpha and beta expression in a case matched series of serous and endometrioid adenocarcinomas of the ovary. Eur J Gynaecol Oncol 29: 126-128, 2008.

178

56. Giuntoli Rn, Rodriguez G, Whitaker R, Dodge R, and Voynow J. Mucin gene expression in ovarian cancers. Cancer Res 58: 5546-5550, 1998.

57. Glish G, and Vachet R. The basics of mass spectrometry in the twenty-first century. Nat Rev Drug Discov 2: 140-150, 2003.

58. Goff B, Mandel L, Muntz H, and Melancon C. Ovarian carcinoma diagnosis. Cancer 89: 2068-2075, 2000.

59. Goldberg J, Piver M, Jishi M, and Blumenson L. Age at onset of ovarian cancer in women with a strong family history of ovarian cancer. Gynecol Oncol 66: 3-9, 1997.

60. Gortzak-Uzan L, Ignatchenko A, Evangelou A, Agochiya M, Brown K, St Onge P, Kireeva I, Schmitt-Ulms G, Brown T, Murphy J, Rosen B, Shaw P, Jurisica I, and Kislinger T. A proteome resource of ovarian cancer ascites: integrated proteomic and bioinformatic analyses to identify putative biomarkers. J Proteome Res 7: 339-351, 2008.

61. Griffiths C. Surgical resection of tumor bulk in the primary treatment of ovarian carcinoma. Natl Cancer Inst Monogr 42: 101-104, 1975.

62. Gu J, Crawley J, Ferrell G, Zhang F, Li W, Esmon N, and Esmon C. Disruption of the endothelial cell protein C receptor gene in mice causes placental thrombosis and early embryonic lethality. J Biol Chem 277: 43335-43343, 2002.

63. Gunawardana C, Kuk C, Smith C, Batruch I, Soosaipillai A, and Diamandis E. Comprehensive Analysis of Conditioned Media from Ovarian Cancer Cell Lines Identifies Novel Candidate Markers of Epithelial Ovarian Cancer. J Proteome Res 2009.

64. Gwinn M, Lee N, Rhodes P, Layde P, and Rubin G. Pregnancy, breast feeding, and oral contraceptives and the risk of epithelial ovarian cancer. J Clin Epidemiol 43: 559-568, 1990.

65. Hayes D, Bast R, Desch C, Fritsche HJ, Kemeny N, Jessup J, Locker G, Macdonald J, Mennel R, Norton L, Ravdin P, Taube S, and Winn R. Tumor marker utility grading system: a framework to evaluate clinical utility of tumor markers. J Natl Cancer Inst 88: 1456-1466, 1996.

66. Heintz A, Odicino F, Maisonneuve P, Beller U, Benedet J, Creasman W, Ngan H, Sideri M, and Pecorelli S. Carcinoma of the ovary. J Epidemiol Biostat 6: 107-138, 2001.

179

67. Hellstrom I, Raycraft J, Hayden-Ledbetter M, Ledbetter J, Schummer M, McIntosh M, Drescher C, Urban N, and Hellstrom K. The HE4 (WFDC2) protein is a biomarker for ovarian carcinoma. Cancer Res 63: 3695-3700, 2003.

68. Helzlsouer K, Bush T, Alberg A, Bass K, Zacur H, and Comstock G. Prospective study of serum CA-125 levels as markers of ovarian cancer. JAMA 269: 1123-1126, 1993.

69. Herbst A. The epidemiology of ovarian carcinoma and the current status of tumor markers to detect disease. Am J Obstet Gynecol 170: 1099-1105; discussion 1105-1097, 1994.

70. Hoffman B, Katsaros D, Scorilas A, Diamandis P, Fracchioli S, Rigault de la Longrais I, Colgan T, Puopolo M, Giardina G, Massobrio M, and Diamandis E. Immunofluorometric quantitation and histochemical localisation of kallikrein 6 protein in ovarian cancer tissue: a new independent unfavourable prognostic biomarker. Br J Cancer 87: 763-771, 2002.

71. Hortin G, Sviridov D, and Anderson N. High-abundance polypeptides of the human plasma proteome comprising the top 4 logs of polypeptide abundance. Clin Chem 54: 1608-1616, 2008.

72. Hough C, Cho K, Zonderman A, Schwartz D, and Morin P. Coordinately up-regulated genes in ovarian cancer. Cancer Res 61: 3869-3876, 2001.

73. Hwang H, Quenneville L, Yaziji H, and Gown A. Wilms tumor gene product: sensitive and contextually specific marker of serous carcinomas of ovarian surface epithelial origin. Appl Immunohistochem Mol Morphol 12: 122-126, 2004.

74. Hynes R. Integrins: versatility, modulation, and signaling in cell adhesion. Cell 69: 11-25, 1992.

75. Ichikawa Y, Nishida M, Suzuki H, Yoshida S, Tsunoda H, Kubo T, Uchida K, and Miwa M. Mutation of K-ras protooncogene is associated with histological subtypes in human mucinous ovarian tumors. Cancer Res 54: 33-35, 1994.

76. Isonishi S. [Biomarker in gynecologic malignancies]. Gan To Kagaku Ryoho 31: 1003-1007, 2004.

77. Jacobs I, and Bast RJ. The CA 125 tumour-associated antigen: a review of the literature. Hum Reprod 4: 1-12, 1989.

180

78. Jacobs I, Skates S, MacDonald N, Menon U, Rosenthal A, Davies A, Woolas R, Jeyarajah A, Sibley K, Lowe D, and Oram D. Screening for ovarian cancer: a pilot randomised controlled trial. Lancet 353: 1207-1210, 1999.

79. Jacobs I, Stabile I, Bridges J, Kemsley P, Reynolds C, Grudzinskas J, and Oram D. Multimodal approach to screening for ovarian cancer. Lancet 1: 268-271, 1988.

80. Jarjanazi H, Savas S, Pabalan N, Dennis J, and Ozcelik H. Biological implications of SNPs in signal peptide domains of human proteins. Proteins 70: 394-403, 2008.

81. Jemal A, Siegel R, Ward E, Hao Y, Xu J, and Thun M. Cancer statistics, 2009. CA Cancer J Clin 59: 225-249, 2009.

82. Jurisicova A, Jurisica I, and Kislinger T. Advances in ovarian cancer proteomics: the quest for biomarkers and improved therapeutic interventions. Expert Rev Proteomics 5: 551-560, 2008.

83. Kabawat S, Bast RJ, Bhan A, Welch W, Knapp R, and Colvin R. Tissue distribution of a coelomic-epithelium-related antigen recognized by the monoclonal antibody OC125. Int J Gynecol Pathol 2: 275-285, 1983.

84. Kanemitsu N, Kato M, Miki T, Komatsu S, Okazaki Y, Hayashizaki Y, and Sakai T. Characterization of the promoter of the murine mac25 gene. Biochem Biophys Res Commun 279: 251-257, 2000.

85. Karlan B, Jones J, Slamon D, and Lagasse L. Glucocorticoids stabilize HER-2/neu messenger RNA in human epithelial ovarian carcinoma cells. Gynecol Oncol 53: 70-77, 1994.

86. Kim J, Skates S, Uede T, Wong K, Schorge J, Feltmate C, Berkowitz R, Cramer D, and Mok S. Osteopontin as a potential diagnostic biomarker for ovarian cancer. JAMA 287: 1671-1679, 2002.

87. Kim S, Bachman N, Nair T, Goldsmith S, Liebert M, Grossman H, Lomax M, and Carey T. Beta 4 integrin transfection of UM-UC-2 (human bladder carcinoma) cells: stable expression of a spontaneous cytoplasmic truncation mutant with rapid loss of clones expressing intact beta 4. Cancer Res 57: 38-42, 1997.

88. Kirchhoff C, Osterhoff C, and Young L. Molecular cloning and characterization of HE1, a major secretory protein of the human epididymis. Biol Reprod 54: 847-856, 1996.

181

89. Kitajiri S, Hosaka N, Hiraumi H, Hirose T, and Ikehara S. Increased expression of integrin beta-4 in papillary thyroid carcinoma with gross lymph node metastasis. Pathol Int 52: 438-441, 2002.

90. Kitteringham N, Jenkins R, Lane C, Elliott V, and Park B. Multiple reaction monitoring for quantitative biomarker analysis in proteomics and metabolomics. J Chromatogr B Analyt Technol Biomed Life Sci 877: 1229-1239, 2009.

91. Klein A, Amigo L, Retamal M, Morales M, Miquel J, Rigotti A, and Zanlungo S. NPC2 is expressed in human and murine liver and secreted into bile: potential implications for body cholesterol homeostasis. Hepatology 43: 126-133, 2006.

92. Komatsu S, Okazaki Y, Tateno M, Kawai J, Konno H, Kusakabe M, Yoshiki A, Muramatsu M, Held W, and Hayashizaki Y. Methylation and downregulated expression of mac25/insulin-like growth factor binding protein-7 is associated with liver tumorigenesis in SV40T/t antigen transgenic mice, screened by restriction landmark genomic scanning for methylation (RLGS-M). Biochem Biophys Res Commun 267: 109-117, 2000.

93. Kuk C, Kulasingam V, Gunawardana C, Smith C, Batruch I, and Diamandis E. Mining the ovarian cancer ascites proteome for potential ovarian cancer biomarkers. Mol Cell Proteomics 8: 661-669, 2009.

94. Kulasingam V, and Diamandis E. Proteomics analysis of conditioned media from three breast cancer cell lines: a mine for biomarkers and therapeutic targets. Mol Cell Proteomics 6: 1997-2011, 2007.

95. Kulasingam V, and Diamandis E. Strategies for discovering novel cancer biomarkers through utilization of emerging technologies. Nat Clin Pract Oncol 5: 588-599, 2008.

96. Kulasingam V, and Diamandis E. Tissue culture-based breast cancer biomarker discovery platform. Int J Cancer 123: 2007-2012, 2008.

97. Kulasingam V, Smith C, Batruch I, Buckler A, Jeffery D, and Diamandis E. "Product ion monitoring" assay for prostate-specific antigen in serum using a linear ion-trap. J Proteome Res 7: 640-647, 2008.

98. Kupryjanczyk J, Thor A, Beauchamp R, Merritt V, Edgerton S, Bell D, and Yandell D. p53 gene mutations and protein accumulation in human ovarian cancer. Proc Natl Acad Sci U S A 90: 4961-4965, 1993.

182

99. Lau S, Sham J, Xie D, Tzang C, Tang D, Ma N, Hu L, Wang Y, Wen J, Xiao G, Zhang W, Lau G, Yang M, and Guan X. Clusterin plays an important role in hepatocellular carcinoma metastasis. Oncogene 25: 1242-1250, 2006.

100. Liotta L, and Kohn E. The microenvironment of the tumour-host interface. Nature 411: 375-379, 2001.

101. Liou H, Dixit S, Xu S, Tint G, Stock A, and Lobel P. NPC2, the protein deficient in Niemann-Pick C2 disease, consists of multiple glycoforms that bind a variety of sterols. J Biol Chem 281: 36710-36723, 2006.

102. Luo L, Soosaipillai A, Grass L, and Diamandis E. Characterization of human kallikreins 6 and 10 in ascites fluid from ovarian cancer patients. Tumour Biol 27: 227-234, 2006.

103. Maines-Bandiera S, and Auersperg N. Increased E-cadherin expression in ovarian surface epithelium: an early step in metaplasia and dysplasia? Int J Gynecol Pathol 16: 250-255, 1997.

104. Mallick P, Schirle M, Chen S, Flory M, Lee H, Martin D, Ranish J, Raught B, Schmitt R, Werner T, Kuster B, and Aebersold R. Computational prediction of proteotypic peptides for quantitative proteomics. Nat Biotechnol 25: 125-131, 2007.

105. Marcotte E. How do shotgun proteomics algorithms identify proteins? Nat Biotechnol 25: 755-757, 2007.

106. Martin J, Eynstone L, Davies M, Williams J, and Steadman R. The role of ADAM 15 in glomerular mesangial cell migration. J Biol Chem 277: 33683-33689, 2002.

107. Martin V. Ovarian cancer. Semin Oncol Nurs 18: 174-183, 2002.

108. Martin V. Straight talk about ovarian cancer. Nursing 35: 36-41; quiz 42, 2005.

109. Maruya S, Myers J, Weber R, Rosenthal D, Lotan R, and El-Naggar A. ICAM-5 (telencephalin) gene expression in head and neck squamous carcinoma tumorigenesis and perineural invasion! Oral Oncol 41: 580-588, 2005.

183

110. Maxfield F, and Menon A. Intracellular sterol transport and distribution. Curr Opin Cell Biol 18: 379-385, 2006.

111. McIntosh M, Liu Y, Drescher C, Urban N, and Diamandis E. Validation and characterization of human kallikrein 11 as a serum marker for diagnosis of ovarian carcinoma. Clin Cancer Res 13: 4422-4428, 2007.

112. Meyer T, and Rustin G. Role of tumour markers in monitoring epithelial ovarian cancer. Br J Cancer 82: 1535-1538, 2000.

113. Moretti R, Marelli M, Mai S, Cariboni A, Scaltriti M, Bettuzzi S, and Limonta P. Clusterin isoforms differentially affect growth and motility of prostate cells: possible implications in prostate tumorigenesis. Cancer Res 67: 10325-10333, 2007.

114. Moss E, Hollingworth J, and Reynolds T. The role of CA125 in clinical practice. J Clin Pathol 58: 308-312, 2005.

115. Murphy G. The ADAMs: signalling scissors in the tumour microenvironment. Nat Rev Cancer 8: 929-941, 2008.

116. Nagele F, Petru E, Medl M, Kainz C, Graf A, and Sevelda P. Preoperative CA 125: an independent prognostic factor in patients with stage I epithelial ovarian cancer. Obstet Gynecol 86: 259-264, 1995.

117. Najy A, Day K, and Day M. ADAM15 supports prostate cancer metastasis by modulating tumor cell-endothelial cell interaction. Cancer Res 68: 1092-1099, 2008.

118. Najy A, Day K, and Day M. The ectodomain shedding of E-cadherin by ADAM15 supports ErbB receptor activation. J Biol Chem 283: 18393-18401, 2008.

119. Nasca P, Greenwald P, Chorost S, Richart R, and Caputo T. An epidemiologic case-control study of ovarian cancer and reproductive factors. Am J Epidemiol 119: 705-713, 1984.

120. Natali P, Nicotra M, Botti C, Mottolese M, Bigotti A, and Segatto O. Changes in expression of alpha 6/beta 4 integrin heterodimer in primary and metastatic breast cancer. Br J Cancer 66: 318-322, 1992.

184

121. Naureckiene S, Sleat D, Lackland H, Fensom A, Vanier M, Wattiaux R, Jadot M, and Lobel P. Identification of HE1 as the second gene of Niemann-Pick C disease. Science 290: 2298-2301, 2000.

122. Ness R, and Cottreau C. Possible role of ovarian epithelial inflammation in ovarian cancer. J Natl Cancer Inst 91: 1459-1467, 1999.

123. Norris H, and Jensen R. Relative frequency of ovarian neoplasms in children and adolescents. Cancer 30: 713-719, 1972.

124. Obata K, Morland S, Watson R, Hitchcock A, Chenevix-Trench G, Thomas E, and Campbell I. Frequent PTEN/MMAC mutations in endometrioid but not serous or mucinous epithelial ovarian tumors. Cancer Res 58: 2095-2097, 1998.

125. Olayioye M, Neve R, Lane H, and Hynes N. The ErbB signaling network: receptor heterodimerization in development and cancer. EMBO J 19: 3159-3167, 2000.

126. Omenn G, States D, Adamski M, Blackwell T, Menon R, Hermjakob H, Apweiler R, Haab B, Simpson R, Eddes J, Kapp E, Moritz R, Chan D, Rai A, Admon A, Aebersold R, Eng J, Hancock W, Hefta S, Meyer H, Paik Y, Yoo J, Ping P, Pounds J, Adkins J, Qian X, Wang R, Wasinger V, Wu C, Zhao X, Zeng R, Archakov A, Tsugita A, Beer I, Pandey A, Pisano M, Andrews P, Tammen H, Speicher D, and Hanash S. Overview of the HUPO Plasma Proteome Project: results from the pilot phase with 35 collaborating laboratories and multiple analytical groups, generating a core dataset of 3020 proteins and a publicly-available database. Proteomics 5: 3226-3245, 2005.

127. Otsuka J, Okuda T, Sekizawa A, Amemiya S, Saito H, Okai T, Kushima M, and Tachikawa T. K-ras mutation may promote carcinogenesis of endometriosis leading to ovarian clear cell carcinoma. Med Electron Microsc 37: 188-192, 2004.

128. Ozols R, Bookman M, Connolly D, Daly M, Godwin A, Schilder R, Xu X, and Hamilton T. Focus on epithelial ovarian cancer. Cancer Cell 5: 19-24, 2004.

129. Palacios J, and Gamallo C. Mutations in the beta-catenin gene (CTNNB1) in endometrioid ovarian carcinomas. Cancer Res 58: 1344-1347, 1998.

130. Park D, Yeo S, Wilson M, Yerbury J, Kwong J, Welch W, Choi Y, Birrer M, Mok S, and Wong K. Clusterin interacts with Paclitaxel and confer Paclitaxel resistance in ovarian cancer. Neoplasia 10: 964-972, 2008.

185

131. Partridge E, and Barnes M. Epithelial ovarian cancer: prevention, diagnosis, and treatment. CA Cancer J Clin 49: 297-320, 1999.

132. Perou C, Sorlie T, Eisen M, van de Rijn M, Jeffrey S, Rees C, Pollack J, Ross D, Johnsen H, Akslen L, Fluge O, Pergamenschikov A, Williams C, Zhu S, Lonning P, Borresen-Dale A, Brown P, and Botstein D. Molecular portraits of human breast tumours. Nature 406: 747-752, 2000.

133. Provencher D, Lounis H, Champoux L, Tetrault M, Manderson E, Wang J, Eydoux P, Savoie R, Tonin P, and Mes-Masson A. Characterization of four novel epithelial ovarian cancer cell lines. In Vitro Cell Dev Biol Anim 36: 357-361, 2000.

134. Redondo M, Villar E, Torres-Munoz J, Tellez T, Morell M, and Petito C. Overexpression of clusterin in human breast carcinoma. Am J Pathol 157: 393-399, 2000.

135. Rhodes DR, Yu J, Shanker K, Deshpande N, Varambally R, Ghosh D, Barrette T, Pandey A, and Chinnaiyan AM. ONCOMINE: a cancer microarray database and integrated data-mining platform. Neoplasia 6: 1-6, 2004.

136. Rifai N, Gillette M, and Carr S. Protein biomarker discovery and validation: the long and uncertain path to clinical utility. Nat Biotechnol 24: 971-983, 2006.

137. Risch H. Hormonal etiology of epithelial ovarian cancer, with a hypothesis concerning the role of androgens and progesterone. J Natl Cancer Inst 90: 1774-1786, 1998.

138. Rizzi F, and Bettuzzi S. Targeting Clusterin in prostate cancer. J Physiol Pharmacol 59 Suppl 9: 265-274, 2008.

139. Robertson D, Pruysers E, and Jobling T. Inhibin as a diagnostic marker for ovarian cancer. Cancer Lett 249: 14-17, 2007.

140. Roland I, Yang W, Yang D, Daly M, Ozols R, Hamilton T, Lynch H, Godwin A, and Xu X. Loss of surface and cyst epithelial basement membranes and preneoplastic morphologic changes in prophylactic oophorectomies. Cancer 98: 2607-2623, 2003.

141. Rose P, Piver M, Tsukada Y, and Lau T. Metastatic patterns in histologic variants of ovarian cancer. An autopsy study. Cancer 64: 1508-1513, 1989.

186

142. Rosen D, Wang L, Atkinson J, Yu Y, Lu K, Diamandis E, Hellstrom I, Mok S, Liu J, and Bast RJ. Potential markers that complement expression of CA125 in epithelial ovarian cancer. Gynecol Oncol 99: 267-277, 2005.

143. Roth L, Emerson R, and Ulbright T. Ovarian endometrioid tumors of low malignant potential: a clinicopathologic study of 30 cases with comparison to well-differentiated endometrioid adenocarcinoma. Am J Surg Pathol 27: 1253-1259, 2003.

144. Sakayori M, Nozawa S, Udagawa Y, Chin K, Lee S, Sakuma T, Iizuka R, Wada Y, Yoshida S, and Takeda Y. [Biological properties of two newly established cell lines (RMUG-S, RMUG-L) from a human ovarian mucinous cystadenocarcinoma]. Hum Cell 3: 52-56, 1990.

145. Sardana G, Jung K, Stephan C, and Diamandis E. Proteomic analysis of conditioned media from the PC3, LNCaP, and 22Rv1 prostate cancer cell lines: discovery and validation of candidate prostate cancer biomarkers. J Proteome Res 7: 3329-3338, 2008.

146. Sato N, Tsunoda H, Nishida M, Morishita Y, Takimoto Y, Kubo T, and Noguchi M. Loss of heterozygosity on 10q23.3 and mutation of the tumor suppressor gene PTEN in benign endometrial cyst of the ovary: possible sequence progression from benign endometrial cyst to endometrioid carcinoma and clear cell carcinoma of the ovary. Cancer Res 60: 7052-7056, 2000.

147. Schildkraut J, Schwingl P, Bastos E, Evanoff A, and Hughes C. Epithelial ovarian cancer risk among women with polycystic ovary syndrome. Obstet Gynecol 88: 554-559, 1996.

148. Shan S, Scorilas A, Katsaros D, Rigault de la Longrais I, Massobrio M, and Diamandis E. Unfavorable prognostic value of human kallikrein 7 quantified by ELISA in ovarian cancer cytosols. Clin Chem 52: 1879-1886, 2006.

149. Shannan B, Seifert M, Leskov K, Willis J, Boothman D, Tilgen W, and Reichrath J. Challenge and promise: roles for clusterin in pathogenesis, progression and therapy of cancer. Cell Death Differ 13: 12-19, 2006.

150. Shih I, Salani R, Fiegl M, Wang T, Soosaipillai A, Marth C, Muller-Holzner E, Gastl G, Zhang Z, and Diamandis E. Ovarian cancer specific kallikrein profile in effusions. Gynecol Oncol 105: 501-507, 2007.

151. Skubitz A, Bast RJ, Wayner E, Letourneau P, and Wilke M. Expression of alpha 6 and beta 4 integrins in serous ovarian carcinoma correlates with expression of the basement membrane protein laminin. Am J Pathol 148: 1445-1461, 1996.

187

152. Sleat D, Wiseman J, El-Banna M, Price S, Verot L, Shen M, Tint G, Vanier M, Walkley S, and Lobel P. Genetic evidence for nonredundant functional cooperativity between NPC1 and NPC2 in lipid transport. Proc Natl Acad Sci U S A 101: 5886-5891, 2004.

153. Slotman B, and Rao B. Ovarian cancer (review). Etiology, diagnosis, prognosis, surgery, radiotherapy, chemotherapy and endocrine therapy. Anticancer Res 8: 417-434, 1988.

154. Soslow R. Histologic subtypes of ovarian carcinoma: an overview. Int J Gynecol Pathol 27: 161-174, 2008.

155. St Croix B, Rago C, Velculescu V, Traverso G, Romans K, Montgomery E, Lal A, Riggins G, Lengauer C, Vogelstein B, and Kinzler K. Genes expressed in human tumor endothelium. Science 289: 1197-1202, 2000.

156. Steen H, and Mann M. The ABC's (and XYZ's) of peptide sequencing. Nat Rev Mol Cell Biol 5: 699-711, 2004.

157. Subramanian A, Sharma A, and Mokbel K. Insulin-like growth factor binding proteins and breast cancer. Breast Cancer Res Treat 107: 181-194, 2008.

158. Sundfeldt K, Piontkewitz Y, Ivarsson K, Nilsson O, Hellberg P, Brannstrom M, Janson P, Enerback S, and Hedin L. E-cadherin expression in human epithelial ovarian cancer and normal ovary. Int J Cancer 74: 275-280, 1997.

159. Suzuki K, and Hayashi T. Protein C and its inhibitor in malignancy. Semin Thromb Hemost 33: 667-672, 2007.

160. Tagliabue E, Ghirelli C, Squicciarini P, Aiello P, Colnaghi M, and Menard S. Prognostic value of alpha 6 beta 4 integrin expression in breast carcinomas is affected by laminin production from tumor cells. Clin Cancer Res 4: 407-410, 1998.

161. Tani T, Karttunen T, Kiviluoto T, Kivilaakso E, Burgeson R, Sipponen P, and Virtanen I. Alpha 6 beta 4 integrin and newly deposited laminin-1 and laminin-5 form the adhesion mechanism of gastric carcinoma. Continuous expression of laminins but not that of collagen VII is preserved in invasive parts of the carcinomas: implications for acquisition of the invading phenotype. Am J Pathol 149: 781-793, 1996.

162. Teneriello M, Ebina M, Linnoila R, Henry M, Nash J, Park R, and Birrer M. p53 and Ki-ras gene mutations in epithelial ovarian neoplasms. Cancer Res 53: 3103-3108, 1993.

188

163. Tian L, Lappalainen J, Autero M, Hanninen S, Rauvala H, and Gahmberg C. Shedded neuronal ICAM-5 suppresses T-cell activation. Blood 111: 3615-3625, 2008.

164. Tonin P, Hudson T, Rodier F, Bossolasco M, Lee P, Novak J, Manderson E, Provencher D, and Mes-Masson A. Microarray analysis of gene expression mirrors the biology of an ovarian cancer model. Oncogene 20: 6617-6626, 2001.

165. Tornos C, Silva E, Ordonez N, Gershenson D, Young R, and Scully R. Endometrioid carcinoma of the ovary with a prominent spindle-cell component, a source of diagnostic confusion. A report of 14 cases. Am J Surg Pathol 19: 1343-1353, 1995.

166. Tuxen M, Soletormos G, and Dombernowsky P. Tumor markers in the management of patients with ovarian cancer. Cancer Treat Rev 21: 215-245, 1995.

167. Vance J. Lipid imbalance in the neurological disorder, Niemann-Pick C disease. FEBS Lett 580: 5518-5524, 2006.

168. Vang R, Gown A, Barry T, Wheeler D, and Ronnett B. Immunohistochemistry for estrogen and progesterone receptors in the distinction of primary and metastatic mucinous tumors in the ovary: an analysis of 124 cases. Mod Pathol 19: 97-105, 2006.

169. Vang R, Gown A, Barry T, Wheeler D, Yemelyanova A, Seidman J, and Ronnett B. Cytokeratins 7 and 20 in primary and secondary mucinous tumors of the ovary: analysis of coordinate immunohistochemical expression profiles and staining distribution in 179 cases. Am J Surg Pathol 30: 1130-1139, 2006.

170. Wang H, Rosen D, Wang H, Fuller G, Zhang W, and Liu J. Insulin-like growth factor-binding protein 2 and 5 are differentially regulated in ovarian cancer of different histologic types. Mod Pathol 19: 1149-1156, 2006.

171. Wang Y, Wu R, Cho K, Shedden K, Barder T, and Lubman D. Classification of cancer cell lines using an automated two-dimensional liquid mapping method with hierarchical clustering techniques. Mol Cell Proteomics 5: 43-52, 2006.

172. Whittemore A. Characteristics relating to ovarian cancer risk: implications for prevention and detection. Gynecol Oncol 55: S15-19, 1994.

189

173. Williams T, Toups K, Saggese D, Kalli K, Cliby W, and Muddiman D. Epithelial ovarian cancer: disease etiology, treatment, detection, and investigational gene, metabolite, and protein biomarkers. J Proteome Res 6: 2936-2962, 2007.

174. Wulfkuhle J, Liotta L, and Petricoin E. Proteomic applications for the early detection of cancer. Nat Rev Cancer 3: 267-275, 2003.

175. Xie M, Motoo Y, Su S, Mouri H, Ohtsubo K, Matsubara F, and Sawabu N. Expression of clusterin in human pancreatic cancer. Pancreas 25: 234-238, 2002.

176. Xu S, Benoff B, Liou H, Lobel P, and Stock A. Structural basis of sterol binding by NPC2, a lysosomal protein deficient in Niemann-Pick type C2 disease. J Biol Chem 282: 23525-23531, 2007.

177. Yin B, Dnistrian A, and Lloyd K. Ovarian cancer antigen CA125 is encoded by the MUC16 mucin gene. Int J Cancer 98: 737-740, 2002.

178. Young R, Prat J, and Scully R. Ovarian endometrioid carcinomas resembling sex cord-stromal tumors. A clinicopathological analysis of 13 cases. Am J Surg Pathol 6: 513-522, 1982.

179. Yu L. The structure and function of Niemann-Pick C1-like 1 protein. Curr Opin Lipidol 19: 263-269, 2008.

180. Zhang L, Ying W, Mao Y, He H, Liu Y, Wang H, Liu F, Wang K, Zhang D, Wang Y, Wu M, Qian X, and Zhao X. Loss of clusterin both in serum and tissue correlates with the tumorigenesis of esophageal squamous cell carcinoma via proteomics approaches. World J Gastroenterol 9: 650-654, 2003.

181. Zhu Y, Wu R, Sangha N, Yoo C, Cho K, Shedden K, Katabuchi H, and Lubman D. Classifications of ovarian cancer tissues by proteomic patterns. Proteomics 6: 5846-5856, 2006.

182. Zurawski VJ, Orjaseter H, Andersen A, and Jellum E. Elevated serum CA 125 levels prior to diagnosis of ovarian neoplasia: relevance for early detection of ovarian cancer. Int J Cancer 42: 677-680, 1988.