matador and the regulation of cyclin e1 in … placental development and placental pathology ......

188
MATADOR AND THE REGULATION OF CYCLIN E1 IN NORMAL HUMAN PLACENTAL DEVELOPMENT AND PLACENTAL PATHOLOGY BY JOCELYN ELAINE RAY A THESIS SUBMITTED IN CONFORMITY WITH THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY GRADUATE DEPARTMENT OF PHYSIOLOGY UNIVERSITY OF TORONTO © COPYRIGHT BY JOCELYN ELAINE RAY, 2010

Upload: buituong

Post on 06-May-2018

219 views

Category:

Documents


0 download

TRANSCRIPT

MATADOR AND THE REGULATION OF CYCLIN E1 IN

NORMAL HUMAN PLACENTAL DEVELOPMENT AND

PLACENTAL PATHOLOGY

BY

JOCELYN ELAINE RAY

A THESIS SUBMITTED IN CONFORMITY WITH THE REQUIREMENTS

FOR THE DEGREE OF DOCTOR OF PHILOSOPHY

GRADUATE DEPARTMENT OF PHYSIOLOGY

UNIVERSITY OF TORONTO

© COPYRIGHT BY JOCELYN ELAINE RAY, 2010

ii

Matador and the Regulation of cyclin E1 in Normal

Human Placental Development and Placental Pathology

Jocelyn Ray, Doctor of Philosophy, 2010

Graduate Department of Physiology, University of Toronto

Toronto, Ontario, Canada

ABSTRACT

Preeclampsia and molar pregnancy are two devastating placental pathologies characterized by an

immature proliferative trophoblast phenotype accompanied by excessive cell death. It is

therefore of paramount importance to study the regulation of cell fate in the placenta, to gain a

further understanding of the mechanisms that contribute to these diseases.

In this dissertation we report that during normal placental development and in preeclampsia,

Matador (Mtd), a pro-apoptotic member of the Bcl-2 family, has a dual function in regulating

trophoblast cell proliferation and death. Importantly, we reveal a novel role of Mtd-L in

promoting cyclin E1 expression and cell cycle progression.

Of clinical importance, we also identify that both cyclin E1 and the CDK inhibitor p27, are

increased in severe early onset preeclampsia. However, the inhibitory function of p27 in this

pathology may be hampered due to its increased phosphorylation at Ser10, resulting in its nuclear

export. Of equal importance, data presented demonstrate that placentae from severe early onset

preeclampsia display a molecular profile distinct from late onset preeclampsia or intrauterine

growth restricted pregnancies.

iii

In the final data chapter we demonstrate that Mtd is highly expressed in molar tissue, where it

localizes to both apoptotic and proliferative cells. Our data suggests that an abundance of Mtd

and cyclin E1 in conjunction with the low level of p27 may contribute to the hyperproliferative

nature of the disorder.

The body of work in this dissertation uncovers novel insights into the regulation of trophoblast

cell fate. Importantly, the impact of Mtd on cyclin E1 to promote G1-S transition is a novel

mechanism found to regulate trophoblast cell proliferation in normal and pathological

placentation. Equally important is our identification of molecular differences between placental

pathologies that may help to differentiate early and late onset preeclampsia, IUGR and molar

pregnancy.

iv

ACKNOWLEDGEMENTS

I would like to express my appreciation to my supervisors, Dr. Isabella Caniggia and Dr. Andrea

Jurisicova for their guidance throughout my PhD. Through this experience I have learned how to

be a better scientist and mentor. Thank you for the scientific discussions and for the lessons on

life. I have learned a great deal that I will keep with me forever.

I would also like to extend my gratitude to the members of my PhD supervisory committee, Drs.

Stephen Matthews, Mingyao Liu, and Lowelle Langille, who gave me valuable advice

throughout my PhD. I am forever grateful.

I would also like to thank Drs. Lye, Brown, Casper, Rogers, and Kingdom who encouraged me

through the last six years, listened to my ideas and helped me to become a better scientist. Thank

you to Bev Bessey and Cindy Todoroff for all the administrative help, support, and kind words.

I would also like to extend my thanks to the scientists with whom I previously worked; Drs. Jay

Wunder, Ben Alman, Carl Ware, and Peter Greer. They deserve the utmost credit for teaching

me the basics, helping to expand my ability to think logically and creatively, and for stimulating

and cultivating my interest in science.

I would especially like to show my appreciation to all the members of the Caniggia, Jurisicova,

and Kingdom labs for their stimulating scientific discussions, technical support, and

encouragement. I would like to thank Dr. Yuan Xu for taking me under her wing in my initial

years, Julia Garcia, and my Caniggia lab sisters, Livia Deda, Antonella Racano, Tara

Sivasubramaniyam and Manpreet Kalkat, who made the six years in the Caniggia lab feel like a

home. I would like to especially thank Livia, my first student and now a great friend and

confidante. She is one of the most compassionate, understanding, and gracious women I know.

Thank you, Antonella, a breath of fresh air, always quick witted and entertaining, and Julia, Tara

and Manpreet for lighting up the lab. I would like to thank Dr. Jacquie Detmar for being

someone I could always turn to for help, for teaching me technical skills and for working with

me to troubleshoot many experiments. I would like to thank Dr. Alicia Tone, for her great

scientific mind and for her amazing friendship. I would like to send a special thanks Dr. Sascha

v

Drewlo for his constant support both scientifically and personally, he was my rock through the

hard times and will remain a best friend and colleague.

Most importantly I would like to thank my family. To my parents who let me fend for myself; I

know you were always there if I needed you. The two of you have been my true motivation.

Lastly I would like to thank my sister who has been an inspiration as a woman of strength. She is

a reminder of what is truly important in life, and has kept me balanced.

vi

CONTRIBUTIONS

The following people have contributed to the generation of data reported in the present thesis.

Introduction: Portions of the introduction have been published in the following form Ray,J.,

Jurisicova,A., and Caniggia,I. (2008). IFPA Trophoblast Research Award Lecture: The

Dynamic Role of Bcl-2 Family Members in Trophoblast Cell Fate. Placenta.

Chapter 3: Yuan Wu contributed to the data presented in Figure 3.5 by performing the JEG-3

cell fractionation experiment. Dr. Julia Garcia created the GFP-hMtd-L doxycycline inducible

cell line used to generate data in Figure 3.8. Dr. Caniggia aided in the explant culture work

performed in chapter 3 and chapter 4. Placental tissue was collected by the research nursing staff

at Mount Sinai Hospital. Data in chapter three have been published in the following form

Ray,J.E., Garcia,J., Jurisicova,A., and Caniggia,I. (2009). Mtd/Bok takes a swing:

proapoptotic Mtd/Bok regulates trophoblast cell proliferation during human placental

development and in preeclampsia. Cell Death and Differentiation.

Chapter 4: Dr. Barbara Cifra contributed to the data presented in Figure 4.10 by performing

western blot and immunofluorescence analysis of phospho-p27 Ser 10 in developmental samples.

Dr. Tullia Todros kindly provided PE and IUGR placentae used in this chapter.

Chapter 5: Dr. Ori Nevo aided in the collection of the molar placentae.

Future directions: Caspase-3 defient mice were obtained from Dr. Tak Mak‟s laboratory, OCI

UHN. Dr. Jurisicova collected the blastocysts and helped to establish the TS cell lines.

Work presented in this thesis was supported by the Genesis Foundation and CIHR.

vii

TABLE OF CONTENTS

ACKNOWLEDGEMENTS ........................................................................................................ iv

CONTRIBUTIONS ...................................................................................................................... vi

TABLE OF CONTENTS ........................................................................................................... vii

List of Tables ............................................................................................................................... xii

List of Figures ............................................................................................................................. xiii

Abbreviations .............................................................................................................................. xv

1 Introduction .............................................................................................................................. 1

1.1 Human Placental Development ....................................................................................... 1

1.1.1 Early placental development ................................................................................... 1

1.1.2 Trophoblast Differentiation and Placental Establishment ..................................... 7

1.1.3 The Mature Placenta ............................................................................................. 10

1.1.4 Oxygen and Human Placentation ......................................................................... 12

1.2 Cell Death and the Regulation of Apoptosis in the Placenta ...................................... 13

1.2.1 Classical Function: Apoptosis and the Intrinsic Pathway .................................... 14

1.2.2 Apoptosis and the Bcl-2 Family in Placentation .................................................. 18

1.2.3 Mtd in Placental Apoptosis ................................................................................... 21

1.3 Regulation of the Cell Cycle in the Placenta ................................................................ 23

1.3.1 The Cell Cycle ....................................................................................................... 23

1.3.2 Cell Cycle Inhibitors ............................................................................................. 26

1.3.3 Regulation of Proliferation and the Cell Cycle in the Placenta ........................... 31

1.3.4 Role of Bcl-2 Family Members in Cell Fate ......................................................... 34

1.4 Placental Pathology ......................................................................................................... 35

viii

1.4.1 Preeclampsia ......................................................................................................... 37

1.4.2 IUGR ..................................................................................................................... 41

1.4.3 Mtd and the Bcl-2 family in preeclampsia ............................................................ 43

1.5 Complete Molar Pregnancy ........................................................................................... 44

1.5.1 Clinical Detection and Classification of Molar Pregnancy ................................. 45

1.5.2 Trophoblast Biology of the Complete Molar Placenta: Morphological

characteristics and Histopathology ...................................................................... 46

1.5.3 Trophoblast Biology of the Complete Molar Placenta: Molecular

Characteristics ...................................................................................................... 47

1.6 Thesis Hypothesis and Objectives ................................................................................. 48

2 Materials and Methods .......................................................................................................... 51

2.1 Placental Tissue Collection ............................................................................................... 51

2.1.1 Placental samples for studies on preeclamptic pathology .................................... 51

2.1.2 Samples for studies on molar twin pathology ....................................................... 53

2.1.3 Samples for laser capture microdissection ........................................................... 53

2.2 First trimester Villous Explant Culture ............................................................................. 55

2.2.1 Mtd antisense knockdown ..................................................................................... 55

2.2.2 TGF treatment .................................................................................................... 55

2.3 Laser Capture Microdissection ......................................................................................... 56

2.4 RNA Analysis ................................................................................................................... 56

2.5 Antibodies ......................................................................................................................... 57

2.6 Western Blot Analysis ...................................................................................................... 58

2.7 Immuno-precipitation ....................................................................................................... 58

2.8 Peroxidase Based Immunohistochemistry ........................................................................ 58

2.9 Immunofluorescence (IF) Staining ................................................................................... 59

2.10 TUNEL (Terminal Deoxynucleotidyl Transferase-dUTP-Nick End Labeling) ............... 60

2.11 Cell Line Culture and Analysis ......................................................................................... 61

ix

2.11.1 SNP (Sodium nitroprusside) treatment ................................................................. 61

2.11.2 Trypan Blue Exclusion Assay ................................................................................ 61

2.11.3 Cell Viability (MTT) .............................................................................................. 61

2.11.4 Cell Fractionation ................................................................................................. 62

2.11.5 Localization of Mtd to Mitochondria .................................................................... 62

2.11.6 siRNA Treatment ................................................................................................... 62

2.11.7 BrdU Incorporation .............................................................................................. 63

2.11.8 TGF treatment .................................................................................................... 63

2.12 Construction of Stable Cell Line Expressing GFP-hMtdL ............................................... 63

2.13 Statistical analysis ............................................................................................................. 63

3 Pro-apoptotic Mtd/Bok Regulates Trophoblast Cell Proliferation during Human

Placental Development and in Preeclampsia ....................................................................... 65

3.1 Abstract ............................................................................................................................ 65

3.2 Introduction ..................................................................................................................... 66

3.3 Results .............................................................................................................................. 67

3.3.1 Mtd expression in proliferating trophoblast cells ................................................. 67

3.3.2 Mtd localizes to villous trophoblast cells in the G1-phase of the cell cycle .......... 69

3.3.3 Mtd expression can occur independently of cell death during early

placentation ........................................................................................................... 72

3.3.4 Mtd-L is the predominant isoform expressed in proliferative trophoblast cells ... 72

3.3.5 Mtd isoforms are differentially localized within proliferative JEG-3 cells .......... 75

3.3.6 SNP-induced apoptosis promotes mitochondrial localization of Mtd in JEG-3

cells ....................................................................................................................... 75

3.3.7 Inhibition of Mtd-L suppresses cyclin E1 expression ........................................... 78

3.4 Discussion ......................................................................................................................... 81

4 Altered trophoblast proliferation in preeclampsia is associated with increased cyclin

E1expression and abnormal regulation of the cell cycle inhibitor p27 ............................. 86

4.1 ABSTRACT ..................................................................................................................... 86

x

4.2 Introduction ..................................................................................................................... 87

4.3 Results .............................................................................................................................. 89

4.3.1 Mtd expression in proliferating trophoblast cells in preeclampsia ...................... 89

4.3.2 Cyclin E1 and the CDK inhibitor p27 show opposing expression during

normal placentation .............................................................................................. 89

4.3.3 Expression of cyclin E1 and p27 is altered in severe early onset preeclamptic

placentae compared to age matched and term controls ....................................... 95

4.3.4 Post translation regulation of p27 is altered in preeclampsia ............................. 98

4.3.5 Regulation of cyclin E1 and p27 are altered in several placental pathologies .... 98

4.3.6 Phosphorylation of p27 is increased in the early stages of normal placental

development ........................................................................................................ 102

4.3.7 TGF influences cyclin E1 and p27 expression in villous explants cultured

under varying oxygen conditions ........................................................................ 102

4.3.8 TGF influences cyclin E1 and p27 expression in JEG-3 choriocarcinoma cell

line cultured under varying oxygen conditions ................................................... 104

4.4 Discussion ....................................................................................................................... 107

5 Dual role for Mtd in trophoblast proliferation and apoptosis in molar pathology ........ 113

5.1 Abstract .......................................................................................................................... 113

5.2 Introduction ................................................................................................................... 114

5.3 Results ............................................................................................................................ 116

5.3.1 Two cases of a twin pregnancy with a complete hydatidiform mole and

coexistent twin fetus ............................................................................................ 116

5.3.2 Second trimester complete molar placentae display increased trophoblast

proliferation and apoptosis ................................................................................. 118

5.3.3 Pro-apoptotic Mtd is elevated in the molar placenta compared to its co-

existing twin and it is associated with apoptotic cells in the trophoblastic and

stromal areas ...................................................................................................... 121

5.3.4 Mtd localizes to the nuclei of proliferative trophoblast cells in molar

pathology ............................................................................................................. 123

5.3.5 Mtd is associated with increased cyclin E1 in villous trophoblast cells of

molar placentae .................................................................................................. 123

xi

5.3.6 Molar placentae exhibit decreased levels of cell cycle inhibitor p27 ................. 127

5.3.7 Molar placentae exhibit altered expression of molecules involved in

regulating the G1 phase of the cell cycle ............................................................ 127

5.4 Discussion ....................................................................................................................... 127

6 Summary and Future Directions ........................................................................................ 136

7 Future Directions – Experimental Design and Preliminary Data ................................... 147

7.1 Determine if caspase-3 is the connecting link between Mtd and cyclin E1 through the

cleavage of p21 and p27. ................................................................................................ 147

7.2 Determine the upstream pathway in preeclampsia leading to the phosphorylation of

p27at Ser10 and its translocation to the cytoplasm ......................................................... 149

7.3 Determine the mechanism leading to low p27 expression in molar tissue and

determine if it contributes to increased Mtd and cyclin E1 expression in the pathology 151

xii

List of Tables

Table 2-1: Clinical data for preeclamptic, intra uterine growth restricted, and control cases ...... 52

Table 2-2: Clinical data for molar twin pregnancies and age matched control twin cases ........... 54

Table 3-1 Expression of Ki67 and Mtd in trophoblast cells ......................................................... 70

xiii

List of Figures

Figure 1-1: Implantation of the blastocyst ...................................................................................... 2

Figure 1-2: Primitive placenta ........................................................................................................ 4

Figure 1-3: Development of placental villi ..................................................................................... 5

Figure 1-4: Placental floating and anchoring villi .......................................................................... 6

Figure 1-5: The placental membrane .............................................................................................. 8

Figure 1-6: The mature placenta ................................................................................................... 11

Figure 1-7: The extrinsic and intrinsic cell death pathway ........................................................... 15

Figure 1-8: BCL-2 family members ............................................................................................. 17

Figure 1-9: Mtd isoforms ............................................................................................................. 22

Figure 1-10: The cell cycle ........................................................................................................... 24

Figure 1-11: Function of cyclin E1 ............................................................................................... 27

Figure 1-12: Regulation and function of p27 ................................................................................ 29

Figure 1-13: Dual role of Bcl-2 family members in cell death and proliferation ......................... 36

Figure 3-1: Mtd expression in proliferating trophoblast cells. ..................................................... 68

Figure 3-2: Association of Mtd with cyclin E1. ........................................................................... 71

Figure 3-3: Apoptosis in early first trimester placental sections. ................................................. 73

Figure 3-4: Mtd isoform mRNA expression in trophoblast subpopulations................................. 74

Figure 3-5: Subcellular localization of Mtd isoforms in JEG-3 cells. .......................................... 76

Figure 5-1 Morphologic characteristics of placentae from the mole and its co-existing twin ... 117

xiv

Figure 5-2 Proliferative assessment of placentae from the mole and its co-existing twin ......... 119

Figure 5-3 Apoptotic assessment of placentae from the mole and its co-existing twin by TUNEL

staining ........................................................................................................................................ 120

Figure 5-4 Expression of apoptotic molecules in molar twins and control twins ....................... 122

Figure 5-5 Co-localization of Mtd with Ki67 expression in mole and twin placentae ............... 124

Figure 5-6 Mtd is expressed in proliferative trophoblast cells associated with various molar

characteristics .............................................................................................................................. 125

Figure 5-7 Cyclin E1 is overexpressed in molar placentae compared to twin controls .............. 126

Figure 5-8 p27 expression in molar placentae and twin controls ............................................... 128

Figure 5-9 Expression of G1 phase cell cycle regulators in the molar and control twins .......... 129

Figure 6-1 Putative model of the mechanism linking Mtd to cyclin E1 expression ................... 139

Figure 7-1 Caspase-3 cleavage of CDK inhibitors in the placenta ............................................. 148

Figure 7-2 Caspase-3 null TS cell derivation ............................................................................. 150

xv

Abbreviations

ABC avidin biotin complex

AMC age matched control

AS antisence

Bcl-2 B cell lymphoma-2

Bok Bcl-2 related ovarian killer

BSA bovine serum albumin

oC degree Celsius

Casp-3 caspase-3

CDK cyclin dependent kinase

CT cytotrophoblast cells

Ct threshold cycle

CTB cytotrophoblast

Cy D/E cyclin D, cyclin E

DAB diaminobenzidine tetraaminobiphenyl

DAPI 4‟.6-diamidino-2-phenylindole

DEPC diethyl pyrocarbonate

dH2O distilled water

DMEM dulbecco‟s modified essential medium

DMSO dimethyl sulfoxide

DNA deoxyribonucleic acid

ECM extracellular matrix

EGF epidermal growth factor

EVT extravillous trophoblast

FBS fetal bovine serum

GTD gestation trophoblastic disease

H2O2 hydrogen peroxide

hCG human chorionic gonadotropin

HIF-1 hypoxia-inducible factor-1

HLA-G histocompatibility-linked antigen-G

HRE hypoxia response element

IF immunofluorescence

INK inhibitor of CDK4

IUGR intra uterine growth restriction

xvi

ug microgram

ul microlitre

Mcl-1 myeloid cell leukemia sequence 1

mg milligram

mL millilitre

mM millimolar

mmHg millimeters of mercury

MMP matrix metalloproteinase

mRNA messenger ribonucleic acid

Mtd matador

MTT 3-[4,5-dimethyl-thiazol-2-yl]-2,5-diphenyltetrazolium bromide

n number of samples

O2 molecular oxygen

OCT optimal cutting temperature

PBS phosphate-buffered saline

PCNA proliferating cell nuclear antigen

PCR polymerase chain reaction

PE preeclampsia

PFA paraformaldehyde

pO2 partial pressure of oxygen

qRT-PCR quantitative real-time PCR

S sense

SS scramble sequence

SEM standard error of the mean

SK syncytial knot

SNP sodium nitroprusside

ST syncytium/syncytiotrophoblast

STBM syncytiotrophoblast microfragments

TBS tris buffered saline

TdT terminal deoxynucleotide transferase

TGF-B transforming growth factor beta

TS trophoblast stem cells (murine)

TUNEL Terminal Deoxynucleotidyl Transferase-dUTP-Nick End Labeling

Ub ubiquitin

VEGF vascular endothelial growth factor

Vol volume

Wks weeks

Wt weight

1

1 Introduction

The placenta is an extremely unique organ that, although transient, has the capacity to sustain

life. During pregnancy it functions as a life line, mediating the physiological exchange between

mother and fetus. Proper formation of the placenta is therefore essential for fetal development

and a successful healthy pregnancy. In humans, cell proliferation, differentiation and death are

the driving forces behind the process of placentation, determining the fate of trophoblast cells.

Abnormality at any stage of this development, due to altered proliferation, differentiation or cell

death may lead to improper placental function and subsequent pregnancy related complications.

It is therefore imperative that mechanisms regulating these cell fate events be elucidated to aid in

prevention, diagnosis, and treatment of placental related disorders.

1.1 Human Placental Development

1.1.1 Early placental development

Human placental development begins with fertilization, a process where the male and female

gametes unite to create a zygote containing a unique set of 46 chromosomes of equal maternal

and paternal contribution. This process, is followed directly by the cortical reaction of the cells

surrounding the ovum to prevent the chance of polyspermy (Moore and Persaud, 1998). In rare

cases gametes of unequal genetic contribution are formed, leading to abnormal pregnancy. This

topic will be discussed further under the context of placental pathologies.

Following fertilization, the resulting zygote travels down the fallopian tube reaching the uterine

cavity within four to five days. Meanwhile the zygote undergoes a number of mitotic cell

divisions in a process referred to as cleavage. Around the fourth or fifth day, fluid penetrates this

mass of cells to form a hollow ball of around 100 cells referred to as a blastocyst. The blastocyst

consists of an inner mass of cells (known as the inner cell mass) that will give rise to the embryo

proper and extraembryonic tissue, and an outer ring of cells that develop to form the trophoblast

of the placenta (Moore and Persaud, 1998) (Figure 1.1).

By the end of the first week the blastocyst adheres to the endometrial epithelium of the uterine

wall, and the process of implantation begins (Figure 1.1). The trophoblast differentiates into two

distinct layers, an inner layer of proliferative multi-potent mononuclear cytotrophoblast cells and

2

Figure 1-1: Implantation of the blastocyst

Approximately one week after fertilization the blastocyst attaches to the uterine wall at its embryonic pole. Syncytial

projections begin to penetrate and invade the maternal endometrium thereby permitting the process of implantation.

Diagram modified from (Moore and Persaud, 1998) with permission.

3

an outer multinucleated layer with no discernable cell boundaries, called the syncytiotrophoblast.

Implantation is initiated by syncytial projections that arise at the embryonic pole of the conceptus

(Figure 1.1). These projections penetrate the endometrium in a process mediated by factors

secreted from the trophoblast, including integrins, matrix metalloproteinases, laminin, and

fibronectin that allow for trophoblast attachment and invasion. In addition, the

syncytiotrophoblast layer also functions to secrete human chorionic gonadotropin (hCG), a

hormone required for the maintenance of pregnancy (Moore and Persaud, 1998). HCG can be

detected as early as the second week after fertilization and is the most commonly used indicator

of pregnancy.

As the blastocyst continues to invade deeper into the endometrium, vacuoles appear in the

syncytium and fuse to form lacunar spaces (Figure 1.2). These lacunae partially contact the

maternal arteries, veins and glands, effectively filling with maternal blood and glandular

secretions from ruptured endometrial capillaries and eroded uterine glands (Figure 1.2). This

produces a nutritional mix (termed embryotroph) that reaches the embryo proper by diffusion

(Moore and Persaud, 1998).

By the end of the second week the underlying cytotrophoblast begins to proliferate and extend

into the syncytiotrophoblast, establishing the formation of the primitive chorionic villi. As the

primitive chorionic villi continue to develop, they begin to branch, expanding into the

intervillous spaces created from the fusion of the syncytial lacunae (Figure 1.3 top panels).

Mesenchyme begins to enter the villous core and soon after capillaries arise, which eventually

fuse to form the arteriocapillary network (Figure 1.3 bottom panel). Meanwhile,

cytotrophoblast cells continue to proliferate and expand out through the syncytium, creating

trophoblast cell columns that form a cytotrophoblastic shell surrounding the chorionic sac, and

effectively anchor the conceptus to the endometrium (Figure 1.3 bottom panel). By the end of

the third week two villous structures have evolved; the anchoring villi that contact the maternal

endometrium, and the floating villi produced from villous branches that remain bathed in the

fluid of the intervillous space (Figure 1.3, 1.4).

4

Figure 1-2: Primitive placenta

The blastocyst invades deeper in to the uterine wall. Cytotrophoblast cells extend into the overlying syncytium to

form the primitive chorionic villi. Lacunar spaces form in the syncytium and accumulate nutritional fluids from

erosion of uterine glands and vasculature that will provide nutrients to the growing conceptus. Diagram modified

from (Moore and Persaud, 1998) with permission.

5

Figure 1-3: Development of placental villi

The chorionic villi begin to branch into the intervillous spaces created from fusion of the syncytial lacunae. In

addition cytotrophoblast cells proliferate and invade past the syncytium to form the trophoblastic shell and establish

the primary extravillous columns and villous tree. Meanwhile mesenchyme and eventually a capillary network

develop in the villous stroma. Diagram modified from (Moore and Persaud, 1998) with permission.

6

Figure 1-4: Placental floating and anchoring villi

The floating villi float in maternal blood within the intervillous space. They are comprised of two trophoblast layers;

an inner layer of proliferative cytotrophoblast cells (green) and an outer layer of differentiated multinucleated

syncytium (white). The core of the villi consists of mesenchymal cells as well as villous vasculature. Anchoring villi

physically attach the placenta to the uterine endothelium (decidua). Anchoring villi are composed of extra villous

trophoblast cells (EVT). The cells at the proximal portion of the column are proliferative (green) and the EVT cells

at the distal end of the column become invasive and migrate into to the deciduas (blue and purple). EVT cells

infiltrate the arterial walls of the maternal spiral arteries and replace cells of the maternal endothelium and smooth

muscle producing uterine arteriole walls consisting of fetal and maternal cells. Green – proliferative trophoblast

cells; blue – differentiating extra villous trophoblast cells; purple – invasive extra villous trophoblast cells.

7

1.1.2 Trophoblast Differentiation and Placental Establishment

1.1.2.1 Formation of the floating villi

The floating villi are comprised of two trophoblast layers, an inner layer of progenitor stem cells,

termed cytotrophoblast, and an outer layer of differentiated syncytium (Figure 1.4). The core of

the villi consists of allantoic mesoderm and the endothelium of the feto-placental circulation. As

the name suggests, these villi float in the maternal fluid, where the branched architecture allow

for maximal surface area for gas and nutrient transport. In early gestation gas and nutrients must

pass through the four layer barrier consisting of syncytium, cytotrophoblast, connective tissue

and endothelium (Figure 1.5).

The layer of syncytium is a ciliated, multinucleated sheath of cellular material with no lateral cell

borders. It is in direct contact with the maternal circulation and acts as a physical barrier between

mother and fetus. The syncytium is also a major source of endocrine activity in the placenta. It

produces and secretes hormones including cortico gonadotropin, placental lactogen, placental

growth hormone and human chorionic gonadotropin (hCG). These secretions are necessary for

placental growth and maternal adaptation to pregnancy.

The syncytial layer is highly differentiated and incapable of self-renewal; a continuous influx of

cellular material is therefore required to keep it replenished so that it can maintain placental

function. This is accomplished through fusion of underlying post-mitotic cytotrophoblast cells

into the syncytial layer, (Huppertz et al., 1998;Potgens et al., 2002;Huppertz et al., 2006). The

addition of cellular material into the syncytiotrophoblast layer is balanced by the deportation of

aged material from the syncytium into the maternal circulation, in the form of membrane

enclosed vesicles termed syncytial knots (Figure 1.4, 1.5). These knots are comprised mostly of

aged nuclei with little cytoplasm. Syncytial knots are carried away through the maternal

circulation and filtered in the maternal lung where they are eventually engulfed by macrophages.

Studies have shown that the process of differentiation from cytotrophoblast to syncytium occurs

in a period of 2-3 days. The material then resides in the syncytium for 3-4 weeks before being

extruded (Huppertz and Kingdom, 2004). The complete process from proliferation through

cytotrophoblast-syncytial fusion, to syncytial shedding is referred to as trophoblast turnover.

Proliferation of cytotrophoblast cells is maximal early in the first trimester in order to support the

initial stages of placental growth, but proliferation is also maintained throughout gestation to

allow for continual trophoblast turnover. Thus trophoblast turnover is highly regulated as

8

Figure 1-5: The placental membrane

The image depicts cross sections through floating placental villi from early gestation and at term. The placental

membrane in the first trimester consists of four layers; the syncytium, the cytotrophoblast cells, the mesenchyme and

the endothelium of the fetal capillaries. At term the membrane is thinned and in some areas consists of only

syncytium and endothelium. Note that at term cytotrophoblast cells persist and are important in replenishing the

syncytial layer that is naturally sloughed off as syncytial knots. Diagram modified from (Moore and Persaud, 1998)

with permission.

9

demonstrated by the time dependent coordination of events. (Specific regulation of trophoblast

cell proliferation and cell death will be covered in more detail in later sections). After the

twentieth week of gestation the placental membrane adapts to the increasing nutritional demands

of the growing fetus. At this point the membrane consists of few cytotrophoblast cells and the

syncytium can become markedly thinned. In addition the capillary endothelium will directly

contact the syncytium in areas, to allow for maximal gas and nutrient transfer (Figure 1.5).

1.1.2.2 Formation of the anchoring villi

The same cytotrophoblast cells that give rise to the syncytial trophoblast also give rise to the

cells that form the anchoring columns and maintain the attachment of the placenta to the uterine

wall. In this case, the majority of the daughter cells produced by the cytotrophoblast, remain as

single non-polarized cells and detach from the basement membrane. As they accumulate, they

form a multilayer of cells that push against the overlying syncytium and break through into the

extravillous space. These cells are referred to as extravillous trophoblast (EVT), since they have

left the confines of the villi (Figure 1.4). The emergence of the EVT constitutes the beginning of

anchoring column formation. Importantly, the cells at the proximal end of the columns maintain

their proliferative potential and function by continuing to feed cells into the column (Figure 1.4).

The EVT cells at the distal portion of the column begin to differentiate, losing their proliferative

capacity, and begin to acquire a migratory phenotype. As the column develops, the EVT cells at

the distal end of the column become invasive and migrate further into to the decidua, up to the

first third of the myometrium (interstitial invasion) (Graham and Lala, 1992) (Figure 1.4).

Differentiation of the extra villous trophoblast cells involves a switch in the expression of a

variety of proteins. As cells differentiate along the invasion pathway they lose cell-cell contact,

and they change their expression pattern of adhesion molecules (Damsky et al., 1994;Caniggia et

al., 1997). Spatially and temporally regulated changes are also seen with respect to the synthesis

and degradation of extracellular matrix proteins and their receptors (Fisher et al., 1989;Caniggia

et al., 1997). The switch in the protein expression accompanying EVT differentiation is governed

in part by a change in the growth factor and cytokine milieu (Morrish et al., 1998). In addition

this process can be opposed by members of the TGF pathway. Specifically, TGF1 and TGF3

have been found to inhibit trophoblast differentiation along the invasive pathway (Graham and

10

Lala, 1991;Caniggia et al., 1999), through a process that is likely mediated by endoglin, a TGF

receptor expressed by the invading EVT (St Jacques et al., 1994;Caniggia et al., 1997).

Trophoblast cells migrate along the capillary walls and essentially plug the maternal spiral

arteries. These endovascular plugs of EVT regress around the 12th week, to allow perfusion of

the inter-villous space by maternal blood. Meanwhile, EVT continue to invade the uterine wall,

and remodel the spiral arteries, to maximize uteroplacental blood flow (Ashton et al., 2005).

Cytotrophoblast cells infiltrate the arterial walls of the maternal spiral arteries and replace cells

of the maternal endothelium and smooth muscle, producing uterine arteriole walls consisting of

fetal and maternal cells (Figure 1.4). This phenomenon has been replicated in vitro using a first

trimester placental-decidual explant model, that showed endothelial cell and smooth muscle

disruption around the maternal vessels, concomitant with trophoblast invasion (Dunk et al.,

2003). This results in the transformation from muscular, narrow, high-resistance uterine

arterioles to non-muscular, distended, low-resistance vessels that are unresponsive to endocrine

stimuli (Kurman, 1991b). This process is critical to the establishment of a successful pregnancy,

as indicated by the various placental pathologies associated with insufficient spiral artery

remodeling (Kurman, 1991a). This is a topic of high importance and will be elaborated upon in

more detail in Section 1.4 of the Introduction.

1.1.3 The Mature Placenta

The mature placenta is thus composed of a fetal chorionic portion and a maternal endometrial

(decidual) component that make up the fetomaternal junction (Figure 1.6). The fetal chorionic

villi anchor the chorionic sac, containing the fetus, to the decidua through the cytotrophoblastic

shell (Moore and Persaud, 1998). Erosion of the decidua, by the extra villous trophoblast,

generates wedge shaped sections of decidua, left behind to form placental septa. These septa

effectively divide the placenta into segmented areas, referred to as cotyledons. Each cotyledon

contains two or more branched stem villi that float in the maternal blood filled intervillous

spaces, created by fusion of the syncytial lacunae (Figure 1.6). Though the intervillous space is

divided by placental septa, maternal blood in the intervillous space can travel freely between

compartments, as the septa do not reach as far down as the chorionic plate. Maternal blood enters

and exits the intervillous spaces through maternal endometrial arteries and veins that pass

through gaps in the trophoblastic shell (Figure 1.6). The maternal blood, rich in oxygen and

11

Figure 1-6: The mature placenta

The mature placenta is a fully functioning organ that mediates gas and nutrient exchange between maternal and fetal

capillary system. Invasive extra villous trophoblast infiltrate the maternal spiral arteries and remodel their

endothelial lining and muscular wall, transforming them into dilated vessels that promote the unencumbered flow of

maternal blood into the maternal space. Floating villi bathe in the maternal blood where their outer covering of

syncytial material performs the exchange of gas and nutrients between fetal and maternal blood. Diagram modified

from (Moore and Persaud, 1998) with permission.

12

nutrients, enters the intervillous space in pulsatile spurts, generated from the maternal blood

pressure. As the blood enters the intervillous space, the pressure lessens and the blood flows

more gently around the branches of placental villi. Gas and nutrients from the maternal blood

diffuse across the placental membrane of the floating villi towards the placental vasculature,

where they are transported to the growing fetus via the fetal blood stream. The fetal capillaries

therefore come in close contact to the maternal blood, but remain separated by the thin placental

membranes.

An intricate artery-capillary-venous return system is established within each chorionic floating

villous. These arteries and veins merge forming larger vessels until they become the two

umbilical arteries and the umbilical vein that connect the placenta to the embryonic heart of the

fetus. Fetal blood begins to flow through this capillary system by the fourth week, allowing

transfer of oxygen and nutrients from the villous space, across the placental villi to the fetal

capillaries, and carbon dioxide and fetal waste are eliminated from the fetus to the maternal

circulation (Moore and Persaud, 1998).

The main function of the placenta is therefore to ensure sufficient nutrient and gas exchange

between the mother and fetus. Additionally the placenta, mediates proper structural attachment

of the conceptus to the uterine wall, is involved in endocrine secretion, provides immunological

protection to the growing fetus, and remodels the maternal spiral arteries to establish a constant,

unhindered blood flow for nutritional purposes.

1.1.4 Oxygen and Human Placentation

As previously mentioned, one of the most important functions of the placenta is to remodel the

maternal vasculature, to allow a sufficient flow of maternal blood into the placental space, so that

oxygen and nutrients can be extracted for embryonic development. However, prior to

establishment of the utero-placental circulation, early embryonic development takes place in a

relatively low oxygenated environment (Burton et al., 1999). It was originally postulated that

oxygenation of the intervillous space occurs in the fourth week of development, following the

initial vascular remodeling events. However, it is now recognized that the placental environment

remains poorly oxygenated well into the late stages of the first trimester, due to obstruction of

the maternal vessels by the formation of the trophoblastic plugs. This environment of low

oxygenation has been well documented as a critical factor leading to proper early placental

13

development, where oxygen itself acts as a key regulator of a variety of cellular events associated

with trophoblast differentiation (Jaffe et al., 1997). The Caniggia laboratory and others have

reported that low oxygen tension can maintain trophoblast cells in a proliferative, non-invasive,

intermediate phenotype, characteristic of early placental development (Genbacev et al.,

1996;Genbacev et al., 1997;Caniggia et al., 2000;MacPhee et al., 2001), and that markers

associated with an immature/intermediate trophoblast phenotype are elevated in a low oxygen

environment (Caniggia et al., 2000;Caniggia and Winter, 2002). In addition, the expression of

hypoxia-inducible factor-1 (HIF-1), a transcription factor that regulates genes in response to low

oxygen, is elevated during this early gestational time period, where it plays a critical role in

mediating the biological effects of oxygen on early trophoblast differentiation (Caniggia et al.,

2000;Rajakumar and Conrad, 2000;MacPhee et al., 2001;Caniggia et al., 2002;Nevo et al.,

2006;Ietta et al., 2007;Yinon et al., 2008). Sufficient remodeling of the spiral arteries and

dislodging of the trophoblastic plugs occurs late in the first trimester, around the 10th

to 12th

week of gestation. This allows for maternal blood to enter the villous space leading to an

increase in oxygenation, which initiates a switch in the expression of a wide variety of genes and

proteins (Rodesch et al., 1992;Burton et al., 1999). Thus, the placenta undergoes a significant

environmental change in the late first trimester. Oxygen electrode studies have shown that prior

to the opening of the intervillous space (around 5-8 weeks of gestation) oxygen tension is around

20 mmHg, (equivalent to 3% O2), whereas at 14-16 wks, after the intervillous space has opened

to maternal blood, oxygen levels rise to ~55-60 mmHg (8-10% O2) (Rodesch et al., 1992;Burton

et al., 1999). In the late stages of pregnancy, the oxygen tension drops slightly closer to ~40

mmHg as a result of increased oxygen extraction from the growing fetus (Soothill et al., 1986).

The rapid increase of oxygen tension in the first trimester has been associated with increased

expression of genes associated with oxidative stress (Jauniaux et al., 2000) and is believed to

drive trophoblast differentiation and death as well as placental maturation (Boyd and Hamilton,

1970;Genbacev et al., 1996;Genbacev et al., 1997;Caniggia et al., 2000). Thus varying degrees

of oxygenation greatly impact upon the behaviour and differentiation of the trophoblast, however

this phenomenon is often overlooked in studies of placental development.

1.2 Cell Death and the Regulation of Apoptosis in the Placenta

Cell death plays a critical role in multiple aspects of placentation, from the attachment of the

blastocyst to the endometrium and its subsequent implantation (Galan et al., 2000), to maternal-

14

placental tolerance (Abrahams et al., 2004), and remodeling of the spiral arteries (Cartwright et

al., 2002;Ashton et al., 2005). Moreover, the development of the placental tissue relies on the

process of cell proliferation, differentiation and death, as they are the driving forces that

determine the fate of each trophoblast cell. Importantly, during placental development cell death

is thought to occur predominantly through the apoptotic pathway. Apoptosis, otherwise referred

to as programmed cell death, is a highly regulated process of cellular self-destruction, employed

to eliminate unnecessary, dysfunctional, or aged cells, so that normal tissue function can be

maintained. As opposed to cell death by way of necrosis, which is controlled by a defined

molecular pathway, however energy independent and often detrimental to the surrounding tissue;

apoptosis is controlled and uses energy dependent processes to facilitate efficient breakdown of

the cell, while avoiding an acute inflammatory response. Cell death by apoptosis is important in

decidualization and remodelling of the spiral arteries. In addition, aspects of the apoptosis

process are essential to the fusion of the trophoblast in the floating villi, and in the deportation of

the syncytial knots. Huppertz et al. have proposed the concept that, during trophoblast turnover,

the apoptotic cascade is initiated in the villous cytotrophoblast, facilitating cell fusion. Apoptosis

is then repressed within the renewed syncytium to maintain cellular function, and then re-

established during the extrusion of apoptotic nuclei in the form of “syncytial knots” (Huppertz et

al., 1998). Therefore, various aspects of the apoptotic pathway contribute to various

differentiation events throughout placentation. A number of studies have assessed late stages of

cell death by TUNEL staining (a marker of the end stages of apoptosis), and found that while the

incidence of trophoblast cell death in chorionic villi is quite low and primarily restricted to the

syncytial layer during the first trimester, the rate of trophoblast cell death significantly increases

with advancing gestation (Smith et al., 1997;Smith et al., 2000). Following a brief introduction to

apoptosis, and the regulation of the intrinsic pathway, the importance and regulation of apoptosis

in placental development will be discussed.

1.2.1 Classical Function: Apoptosis and the Intrinsic Pathway

Two distinct but converging pathways lead to apoptosis; the extrinsic pathway initiated by

activation of cell death receptors located at the plasma membrane, and the intrinsic pathway

mediated by internal cues reflecting cellular homeostasis that target the integrity of the

mitochondrial membrane (Figure 1.7). Both pathways, governed either by external or internal

15

Figure 1-7: The extrinsic and intrinsic cell death pathway

Signalling of apoptosis through the extrinsic and intrinsic pathways is illustrated. The extrinsic pathway is initiated

by cell death receptors located at the plasma membrane which activate initiator caspases (caspase 2,8,9,10). This

results in the cleavage and activation of downstream effector caspases (caspase 3,6,7) leading to apoptosis. The

intrinsic pathway is governed by pro-apoptotic and anti-apoptotic members of the Bcl-2 family. Internal cues

reflecting cellular homeostasis induce or suppress their expression. Elevated levels of pro-apoptotic molecules (Mtd,

Bax, Bak, tBid are depicted) associate and form pores in the mitochondrial membrane leading to the release of

apoptogenic factors including cytochrome c into the cytoplasm. This event leads to the formation of the apoptosome

and activation of effector caspases. The extrinsic and intrinsic pathway converge following activation of effector

caspases or through caspase 8 activation of tBid. The effect of pro-apoptotic Bcl-2 family members can be

suppressed by their interaction with anti-apoptotic Bcl-2 family members (Mcl-1 is depicted).

16

signals, converge to complete a series of events that define apoptosis. Hallmarks of apoptosis

include nuclear condensation, DNA fragmentation, blebbing of the plasma membrane, and cell

shrinkage into dense apoptotic bodies (Straszewski-Chavez et al., 2005). These cellular changes

result largely from the activation of “executioner” or “effector” caspases, a family of cysteine-

aspartate proteases that cleave specific consensus sites on selected target proteins, to facilitate the

characteristic events of apoptosis.

The caspase family can be divided into initiator (caspases -2,8,9,10) and effector (caspases -

3,6,7) members (Degterev et al., 2003), each of which is synthesized as an inactive proenzyme

that is activated upon its cleavage. Initiator caspases are activated by an allosteric mechanism,

facilitated by dimerization or oligomerization. Once activated these members function primarily

in the cleavage and activation of downstream effector caspases, thereby initiating what is known

as the caspase cascade (Degterev et al., 2003;Fuentes-Prior and Salvesen, 2004) (Figure 1.7).

Effector caspases then cleave further procaspases, creating a feed forward cycle, as well as

targeting a variety of vital cellular proteins (Timmer and Salvesen, 2007), resulting in cell

demise.

In addition to caspases, the most widely studied molecules involved in the intrinsic pathway of

apoptosis are members of the bcl-2 family (Figure 1.8). Members of the Bcl-2 family function

predominantly to regulate the integrity of the mitochondrial membrane. This family includes

both pro-apoptotic and anti-apoptotic members that function to either promote the formation of

pores in the mitochondrial membrane, or to prevent pore-formation, respectively. These family

members are identified by the presence of a Bcl-2 homology (BH) domain, of which there are

four (BH1-4). More than 20 members of this family have now been recognized (Gross et al.,

1999), each member containing at least one BH domain (Figure 1.8). These family members are

subdivided in to three groups based on their function and number of BH domains. The anti-

apoptotic or „cell death suppressors‟ (Bcl-2, Bcl-xL, Mcl-1, A1) contain four BH domains

whereas the cell death inducers contain either multiple BH domains (Bax, Bak, and Mtd/Bok) or

a single (BH3-only) domain (Hrk, Bim, Bad, Bik). Importantly, it is the BH3 domain that allows

these family members to interact with one another. Members of this gene family act through a

complex network of promiscuous homo- and hetero-dimers with the exception of Mtd, which

interacts almost exclusively with Mcl-1 (Hsu et al., 1997).

17

Figure 1-8: BCL-2 family members

The three categories of Bcl-2 family members represented by the more well known family members are depicted.

The anti-apoptotic or „cell death suppressors‟ (Bcl-2, Bcl-xL, Bcl-W, Mcl-1, A1) contain four Bcl-2 homology

regions (BH1, purple; BH2, dark blue; BH3, green; BH4, light blue) whereas the cell death inducers contain either

multiple BH domains (Bax, Bak, and Mtd/Bok) or a single (BH3-only) domain (Bid, Bad, Bik, Bim, BNip3, Nix).

Importantly, it is the BH3 domain that allows these family members to interact with one another. In addition many

of the family members possess a hydrophobic carboxy-terminal transmembrane domain (TM, pink) that allows them

to bind to intracellular membranes.

18

The anti-apoptotic members localize to the mitochondrial outer membrane where they protect

against pore formation and leakage of apoptotic inducing factors. Activation of the pro-apoptotic

members, which contain a pore-forming region, are believed to target the mitochondria, causing

release of apoptotic factors which, in turn, results in the activation of the caspase cascade

(Figure 1.7). In contrast, the BH3-only proteins confer their pro-apoptotic function by activating

pro-apoptotic Bax and Bak directly or via neutralizing anti-apoptotic family members. The

balance between pro- and anti-apoptotic molecules thus regulates cell death by controlling the

permeability of the mitochondrial membrane (Figure 1.7).

Depolarization of the mitochondrial membrane results in the release of apoptogenic factors

including cytochrome c (Yang et al., 1997;Brunelle and Chandel, 2002), Smac/Diablo (Du et al.,

2000), and apoptosis inducing factor (AIF) (Susin et al., 1996). Together, with ATP and

apoptosis inducing factor-1 (Apaf-1), these molecules aid in the recruitment of initiator caspase 9

and the formation of the “apoptosome”. Dimerization and allosteric activation of caspase-9

occurs at the apoptosome and consequently leads to caspase-9 mediated activation of the effector

caspases 3, 6 and 7. Protease function of the effector caspases results in cleavage of nuclear

lamins, DNA repair enzymes, and cytoskeletal proteins leading to the events culminating in

cellular apoptosis (Figure 1.7).

1.2.2 Apoptosis and the Bcl-2 Family in Placentation

As mentioned earlier, a variety of processes necessary for proper placentation involve aspects of

apoptosis. Implantation, maternal immune tolerance, and spiral artery remodeling require

trophoblast mediated apoptosis of surrounding cells of the endometrium and decidua (Cartwright

et al., 2002;Ashton et al., 2005). These events are thought to result from trophoblast cell

expression of Fas ligand (FasL), a peptide that interacts with the Fas receptor to stimulate

apoptosis through the extrinsic pathway. The Fas receptor is expressed by endothelial cells of the

endometrium and the villous vessels, as well as maternal T lymphocytes, resulting in trophoblast

mediated apoptosis of these cell types when the two come into contact (Ashton et al., 2005).

Studies conducted in vitro have shown through use of a co-culture system that unmodified (not

placental bed) spiral arteries obtained from cesarean sections co-cultured with extra villous

trophoblast cells exhibited a loss of the endothelial layer. This was associated with signs of

caspase cleaved PARP (poly ADP-ribose polymerase), an apoptosis marker, and could be

19

prevented by pretreatment with inhibitors of caspase or FasL (Ashton et al., 2005;Cartwright and

Wareing, 2006). Destruction of these cells allows for trophoblast migration into the uterine wall

and spiral arteries and inhibits an immune response from maternal leukocytes so that pregnancy

can continue unscathed.

In contrast, apoptotic events that take place within the placenta are thought to occur through a

process that involves regulators of the intrinsic cell death machinery (Huppertz et al.,

1998;Potgens et al., 2002). As previously mentioned, the process of trophoblast turnover is

believed to involve a multi-step cascade of apoptotic events that begin as the cytotrophoblast

prepares for fusion with the overlying syncytium, and is temporarily repressed within the

syncytial layer before its completion during syncytial knot formation and deportation (Huppertz

et al., 1998). Furthermore, it has been determined that members of the Bcl-2 and caspase family

are key mediators of these events (Huppertz et al., 1998;Ray et al., 2008;Heazell and Crocker,

2008a).

Trophoblast cell fate in the floating villi appears to be regulated at different stages by various

members of the Bcl-2 or caspase family. The initial stages of apoptosis occur in the

cytotrophoblast layer, as demonstrated by the expression of the initiator caspase 8 and cleavage

of a-fodrin (a cytoskeleton protein), as well as by the externalization of phosphatidylserine (an

aminophospholipid) from the inner to the outer leaflet of the plasma membrane (Huppertz et al.,

1998;Huppertz et al., 1999). Although these events are markers of early apoptosis, their

expression at this point may play more of a role in trophoblast differentiation then in cell death.

Studies indicate that caspase 8 activity and phosphatidylserine externalization are important for

the fusion of cytotrophoblast cells to the syncytium, as inhibition of either phosphatidylserine or

caspase-8 activation (Black et al., 2004) reduces syncytial fusion in vitro. Other apoptotic

molecules have also been found to be expressed in the cytotrophoblast layer including p53, Bax

and Mtd which may contribute to the process of trophoblast differentiation (Ratts et al.,

2000;Soleymanlou et al., 2005b).

Once the cellular material enters the syncytium further progression of the caspase cascade is

prevented until the extrusion period. Cytoplasmic expression of Mcl-1 in the cytotrophoblast and

syncytium (Huppertz et al., 1998), and Bcl-2 expression primarily in the syncytiotrophoblast

(Ratts et al., 2000;Axt-Fliedner et al., 2001;Danihel et al., 2002), have been postulated to protect

20

against final execution stages of apoptosis (Huppertz et al., 1998;Danihel et al., 2002).

Consistent with this hypothesis is the finding that the expression of Bcl-2 and Mcl-1 is absent

from syncytial knots displaying TUNEL positivity (Huppertz et al., 1998), and the observation

that both Mcl-1 and Bcl-2 levels are elevated during early placentation when trophoblast

apoptosis is minimal (Huppertz et al., 1998).

As the syncytial material ages, portions prepare for the final stages of apoptosis and deportation

of the apoptotic material in the form of membrane sealed fragments termed syncytial knots,

marking the final stages of trophoblast apoptosis (Yasuda et al., 1995;Huppertz et al., 1998).

Cytoplasmic expression of pro-apoptotic Bax and Bak has been found in discrete areas of the

syncytium primarily around regions associated with syncytial knot formation and fibrin

deposition (Ratts et al., 2000). Levels of Bax and Bak, however, have been found to be

consistently low throughout gestation suggesting that they may not be the principal molecules

driving apoptosis in the human placenta (Ratts et al., 2000;De Falco et al., 2001). As pregnancy

progresses the incidence of trophoblast apoptosis increases and by the third trimester as much as

3 grams of syncytial material is shed into the maternal circulation a day (Huppertz et al., 1998).

Efficient clearing of the foreign material, by the maternal immune system, is a natural part of

pregnancy that prevents systemic endothelial cell activation and maintains maternal health

(Redman and Sargent, 2003). However, in some patients the maternal immune system is not

capable of coping with the syncytial debris produced. This can lead to the placental pathology of

preeclampsia (discussed in Introduction 1.4). Furthermore, it has been proposed that in severe

placental pathologies the mode of cell death may change from one of apoptosis towards that of

necrosis, a process referred to as aponecrosis (Huppertz et al., 2003).

Oxygen status has also been shown to influence cell fate of the trophoblast. A low oxygen

environment, such as that experienced in the early first trimester, is associated with increased

trophoblast proliferation and retention of trophoblast cells in the cytotrophoblast layer due to a

lack of trophoblast fusion (Huppertz et al., 2003). In cases of severe hypoxia, or oxidative stress,

poor ATP production may push the regulated process of apoptosis to that of aponecrosis

(Huppertz et al., 2003). Hypoxia has also been shown to increase Bax expression and lower Bcl-

2 expression in cultured first trimester cytotrophoblast cells in some studies, whereas other

studies have found increased p53 and mdm2 under hypoxic conditions, but no alteration in the

expression of either Bax or Bcl-2 (Hu et al., 2006;Heazell et al., 2008b). The Caniggia laboratory

21

has shown that Mcl-1 increases in physiological low oxygen conditions (3% O2) and that this

molecule can protect against Mtd-mediated apoptosis (Soleymanlou et al., 2007). Under

conditions of oxidative stress however, Mcl-1 is cleaved to a pro-apoptotic molecule

(Soleymanlou et al., 2007), underscoring the importance of the oxygen status in the human

placenta. Furthermore, the Caniggia lab has identified a novel Mtd splice variant, abundantly

expressed by the human placenta (Soleymanlou et al., 2005b;Soleymanlou et al., 2007), and

shown that the balance between the anti-apoptotic Mcl-1 and the pro-apoptotic Mtd, is dependent

upon oxygen, and is pivotal in determining placental homeostasis (Soleymanlou et al., 2007).

1.2.3 Mtd in Placental Apoptosis

Mtd/Bok (Mtd: Matador/Bok: Bcl-2-related ovarian killer) is a pro-apoptotic multi-domain pore-

forming member of the Bcl-2 family (Hsu et al., 1997). Mtd mRNA is alternatively spliced and

encodes for three protein isoforms Mtd-L, Mtd-S, and a third placental specific isoform Mtd-P

(Figure 1.9), which has been characterized in normal placental development, as well as in

placental tissue from pregnancies complicated by preeclampsia (Hsu et al., 1997;Soleymanlou et

al., 2005b). Unlike Bax and Bak, Mtd-L and Mtd-P are expressed at very high levels in the

placenta and in tissues of reproductive origin (Hsu et al., 1997;Soleymanlou et al., 2005b)

suggesting that Mtd may be the primary pro-apoptotic Bcl-2 family member regulating

trophoblast apoptosis. Soleymanlou et al previously reported that Mtd is localized to the

syncytial knots in first trimester of human placentae where it is associated with trophoblast

apoptosis and localized to the cytotrophoblast layer and extra villous trophoblast cells where it

may have a regulatory role in trophoblast differentiation (Soleymanlou et al., 2005b).

Similarly to Bax, all isoforms of Mtd contain three BH domains and a transmembrane domain

which are believed to facilitate pro-apoptotic activity via mitochondrial depolarization

(Soleymanlou et al., 2005b) (Figure 1.9). This has been further supported by the finding that

overexpression of Mtd-L or Mtd-P in CHO cells leads to mitochondrial depolarization and

subsequent cleavage and activation of the caspase cascade (Soleymanlou et al., 2005b). It is

suggested, that due to disruption in the BH3 domain of Mtd-S and Mtd-P that occurs as a

consequence of splicing, neither isoform can interact with other Bcl-2 family members. It is

therefore likely that Mtd-S and Mtd-P exert their apoptotic function primarily by forming pores

directly in the mitochondrial membrane, facilitating release of various apoptogenic proteins,

22

Figure 1-9: Mtd isoforms

The three isoforms of Matador (Mtd) are illustrated. Top panel depicts the human chromosomal structure and

transcript maps of the Mtd isoforms. The full length isoform (Mtd long, Mtd-L) consists of five exons. Mtd-S is a

short isoform resulting from exon 3 skipping, whereas the placental specific isoform (Mtd-P) arises from exon 2

skipping. Figure adapted by permission from Macmillan Publishers Ltd: Cell Death and Differentiation,

Soleymanlou et al, copyright 2005. The protein domains are depicted in the lower panel. Of note, all three isoforms

maintain the pore forming region encoded by exon 4. Only Mtd-L maintains an intact BH3 domain required for

member to member interaction.

23

whereas Mtd-L likely mediates its apoptotic activity by antagonizing the function of the anti-

apoptotic Mcl-1, in addition to targeting the mitochondria (Hsu et al., 1997;Soleymanlou et al.,

2005b). Similarly to Mcl-1, low oxygen tension stimulates the expression of Mtd-L and Mtd-P,

which may explain their increased expression in the early stages of placentation and their

increased expression in placental pathologies. As previously mentioned, Mtd expression is high

during early placental development (Soleymanlou et al., 2005b) and is localized to the dynamic

cytotrophoblast and extra villous trophoblast layers. As this gestational period and site of

expression are characterized by intense trophoblast cell proliferation and little trophoblast cell

death, it is likely that Mtd, in addition to its classical role in apoptosis, may have a function in

regulating other areas of trophoblast cell fate, such as trophoblast proliferation.

1.3 Regulation of the Cell Cycle in the Placenta

1.3.1 The Cell Cycle

Adequate proliferation and differentiation of the trophoblast cells is necessary to establish

successful growth of the placenta. Proliferation includes a specific set of events, coordinated

precisely, to progress the cell cycle forward through G1, S, G2 and M phases (Figure 1.10).

During the G1 phase the cell senses and responds to environmental cues prompting the cell to

begin cycling, it then grows duplicating its organelles and preparing for DNA synthesis. During

this initial phase the cell passes a restriction point, after which mitogen stimulation is no longer

required for the cell to continue cycling. This is effectively a point of no return whereby the cell,

from this point onward, is committed to completion of cell division, or will undergo cell death

(Sherr, 1994). At the end of the G1 phase the cell passes the G1/S checkpoint and enters the S

phase where the cell undergoes DNA synthesis. This is followed by progression to the G2 phase

where the integrity of the replicated DNA is assessed before crossing the G2/M checkpoint.

Once in M phase the chromosomes segregate and the cell physically divides. The two most

important decision points therefore occur during the G1 phase: 1) the decision to begin the

cycling process and 2) the decision to commit to cell division (Sherr, 1994). Not surprisingly,

control of the cell cycle occurs largely during the G1 phase, and has subsequently become a

classic target in malignant transformation.

As shown in Figure 1.10, the transition through each stage of a typical mammalian cell cycle is

governed by cyclin dependent kinases (CDKs) that are regulated by specific cyclins (activators)

24

Figure 1-10: The cell cycle

A schematic of the cell cycle is depicted. Mitogenic cues stimulate cyclin D expression to initiate entry into the G1

phase. The D type cyclins then activate cyclin dependent kinases 4 and 6 (CDK4/6) which phosphorylate Rb. Rb is

further phosphorylated by cyclin E activated CDK2. Phosphorylation of Rb prevents its interaction with the E2F

transcription factor leading to E2F dependent gene transcription and entry in to the S phase. Cyclin A in complex

with CDK2 drives cells through the S phase in conjunction with PCNA. Entry into the G2 phase is associated with

cyclin A-CDK1 interaction which is followed by a switch to cyclin B- CDK1 interaction as the cell transitions from

the G2 to M phase. Interaction between cyclin B and CDK1 is then maintained and drives cells through the M phase.

The cell cycle is inhibited by INK4 (p15, p16, p18, and p19) and the Cip/Kip (p21, p27, and p57) family. INK4

(inhibitor of CDK4) family members bind CDK4 and CDK6 to inhibit the G1-S transition by preventing their

association with D type cyclins, while Cip/Kip proteins bind a variety of cyclins, CDKs or cyclin bound CDKs, and

can inhibit the cell cycle at various points.

25

and CDK inhibitors (CDKIs: repressors). Expression of the CDKs remains relatively constant

over the cell cycle period, whereas cyclin expression is phase specific (Sherr, 1994). The phase

sensitivity of the cyclins, in combination with their short half life (approximately 10-25 minutes),

assures that progression of the cell cycle is tightly controlled (Sherr, 1994). Entry into the G1

phase is accompanied by the expression of D type cyclins (D1, D2, D3) known as growth factor

sensing cyclins (Figure 1.10). Their expression is initiated in response to mitogen stimuli,

effectively linking the activation of the cell cycle with environmental cues (Sherr, 1995).

Independently, each D type cyclin can bind to and allosterically regulate one of two CDK

subunits, CDK4 or CDK6 (Morgan, 1997). Assembly of D cyclins into cyclin D-CDK(4/6)

complexes occurs in the cytoplasm through a process mediated by the interaction with a Cip/Kip

protein (Labaer et al., 1997). This interaction is believed to stabilize the complex and aid in its

translocation from the cytoplasm to the nucleus, as neither cyclin D nor CDK4/6 contains its own

nuclear translocation sequence (Labaer et al., 1997;Cheng et al., 1999). Once nuclear the kinase

function of the CDK is activated, and phosphorylation of its target proteins begins. Shortly

thereafter, cyclin E (E1 or E2) is expressed and forms complexes with CDK2, to promote the G1

to S transition (Dulic et al., 1992;Koff et al., 1992) (Figure 1.10).

Both cyclin D and cyclin E activated CDKs target the retinoblastoma family of proteins (pRb,

p107 and p130) which function to inhibit the cell cycle by binding to members of the E2F family

(Dyson, 1998). The E2F family is a group of transcription factors that allow for transcription of

genes important in G1-S transition and required for DNA synthesis (Dyson, 1998;Nevins, 1998).

Phosphorylation of Rb is a multi-step process that is initiated by the cyclin D dependent kinases

and completed by cyclin E-CDK2 complexes (Sherr, 1994;Lundberg and Weinberg, 1998). Once

sufficiently phosphorylated, Rb is released from E2F which allows for E2F-dependent gene

transcription (Adams, 2001) and entry in to the S phase (Figure 1.10).

Importantly, E2F initiates transcription of cyclin E, resulting in a positive feedback loop and

ensuring the irreversible commitment to passage through the G1-S checkpoint (Knoblich et al.,

1994;Geng et al., 1996). Cyclins A and B become important in later stages of the cycle by

promoting G1/S (cyclin A) and G2/M transition (cyclins A and B) (Bailly et al., 1992;Sherr,

1996). Cyclin A in complex with CDK2 drives cells through the S phase in conjunction with

PCNA (proliferating cell nuclear antigen), a protein involved in DNA replication and widely

used as a marker of cell proliferation (Figure 1.10). Entry into the G2 phase is associated with

26

cyclin A-CDK1 interaction which is followed by a switch to cyclin B- CDK1 interaction, as the

cell transitions from the G2 to M phase. Interaction between cyclin B and CDK1 is then

maintained and drives cells through the M phase (Figure 1.10) (Sherr, 1996). Additionally,

Ki67, a protein of unknown function, is expressed throughout the active phases of the mitotic

cell cycle, and has become a widely used marker of proliferation (Endl and Gerdes, 2000).

Although D and E cyclins both target the retinoblastoma family, they have many different

properties. Whereas, specific cell types often express different D-type cyclins and either CDK4

or 6, cyclin E1,E2 and CDK2 expression is not cell type specific (Geng et al., 2001).

Furthermore, unlike the mitogen mediated expression of D type cyclins, expression of cyclin E is

controlled by an autonomous mechanism and peaks sharply at the G1/S border (Dulic et al.,

1992;Koff et al., 1992). Additionally, cyclin E-CDK2 complexes have wider substrate

specificities than the cyclin D dependent kinases that primarily target Rb. These include

phosphorylation of Rb, histone H1, p27 and a variety of other proteins (Sherr and Roberts,

2004;Moroy and Geisen, 2004). Therefore, once activated, cyclin E-CDK2 complexes not only

complete the deactivation of Rb to initiate E2F regulated gene transcription, but they also initiate

a series of additional processes including DNA replication, centrosome duplication, and histone

biosynthesis. (Sherr et al., 2004;Moroy et al., 2004) (Figure 1.11)

Cyclins D and E have also been found to function in processes that are independent of their

catalytic functions. For example, cyclin D-CDK complexes also bind and sequester CDK

inhibitors, freeing the cyclin E-CDK2 complexes from their inhibition (Sherr and Roberts, 1995),

and cyclin E has been found to be required for the assembly of the pre-initiation complex at

origins of DNA replication in quiescent cells entering the cell cycle (Coverley et al., 2002) and in

the process of endoreduplication (MacAuley et al., 1998;Parisi et al., 2003) (Figure 1.11).

1.3.2 Cell Cycle Inhibitors

In mammalian cells, the cell cycle is inhibited by two groups of CDK inhibitors, the INK4 (p15,

p16, p18, and p19) and the Cip/Kip (p21, p27, and p57) family. INK4 (inhibitor of CDK4)

family members bind CDK4 and CDK6 to inhibit the G1-S transition by preventing their

association with D type cyclins, while Cip/Kip proteins bind a variety of cyclins, CDKs or cyclin

bound CDKs, and can inhibit the cell cycle at various points (Sherr and Roberts, 1999) (Figure

1.10).

27

Figure 1-11: Function of cyclin E1

A schematic of the CDK2 dependent and independent functions of cyclin E1. Cyclin E1 activates CDK2

phosphorylation of Rb, CDC25A phosphatase, and p27. This leads to G/S phase transition and a feed forward loop.

The cyclin E1 activated kinase activity of CDK2 is also involved in DNA replication, centrosome duplication, and

histone biosynthesis. CDK2 independent functions of cyclin E1 include roles in endoreduplication, re-entry from

G0-G1, and malignant transformation.

28

P21, originally discovered as a CDK inhibitor and as a mediator of p53 tumor suppression, is

now known to be involved in numerous roles geared toward regulating cell viability, including

cell cycle arrest in response to DNA damage, promotion of differentiation and enforcement of

cellular senescence. The cell cycle arrest function of p21 is reliant on its nuclear expression

where it binds to CDKs. Additionally p21 can bind and prevent the function of PCNA, a protein

required for DNA synthesis, thereby inhibiting cell cycle progression (Sherr et al., 1999).

In addition to its anti-proliferative functions, p21 can also act in a cell cycle promoting manner.

p21 mediates cyclin D-CDK assembly and aids in the transport the complex from the cytoplasm

into the nucleus (Labaer et al., 1997;Cheng et al., 1999;Sherr et al., 1999). The incorporation of

p21 into the cyclinD-CDK4/6 complex does not prevent the kinase activity, but instead further

stabilizes the cyclin-CDK complex within the nucleus, preventing its translocation back to the

cytoplasm, and thus promoting the progression through the G1 phase. Additionally, p21 can also

promote cell viability through an anti-apoptotic mechanism which is achieved by the binding of

p21 to caspase-3 and preventing its activation (Suzuki et al., 1998;Asada et al., 1999).

The p21 related protein, p27 was originally discovered as the mediator of G1 arrest induced by

TGF- or contact inhibition (Polyak et al., 1994). Although p27 can bind and inhibit a number of

CDKs, it was found to preferentially bind to cyclin E/CDK2, inhibiting the CDK2 activity by

interfering with the catalytic cleft and preventing ATP binding (Russo et al., 1996). In addition,

p27 has been linked to the maintenance of cell quiescence associated with mitogen starvation,

and in a number of cell types it is down-regulated upon growth factor stimulation (Sgambato et

al., 2000;Besson et al., 2006). Like p21, p27 can also aid in complex formation of D-type cyclins

with CDK4/6 (Sgambato et al., 2000). This interaction functionally sequesters the Cip/Kip

inhibitors, and subsequently frees the cyclin E-CDK2 complexes from their inhibition, allowing

CDK2 activity to occur. In contrast mitogen withdrawal leads to decreased cyclin D expression,

thereby freeing p21 and p27, enabling them to inhibit cyclin E-CDK2 and arrest the cell cycle

(Sherr et al., 1999). Additional functions of p27 continue to be discovered. Over the past few

years studies have revealed a role for p27 in cell differentiation, cell migration, transcriptional

regulation and apoptosis (Besson et al., 2008) (Figure 1.12).

29

Figure 1-12: Regulation and function of p27

The function and stability of p27 is regulated by subcellular localization, protein interaction and phosphorylation

status. p27 interacts with CDKs in the nucleus to prevent cell cycle progression. P27 is exported to the cytoplasm

following Ser10 phosphorylation and interaction with the exportin CRM1. In the cytoplasm p27 can interact with D

type cyclins and CDKs and aid in cyclin D-CDK assembly. Alternatively p27 can function in cell migration through

its interaction with RhoA. Phosphorylation of p27 at alternative sites can lead to either its stability or degradation by

promoting the interaction of p27 with various protein complexes. Known phosphorylation sites are depicted by the

colored circles.

30

1.3.2.1 Regulation by Phosphorylation

The function and stability of p21 and p27 is highly regulated by subcellular localization, protein

interaction and phosphorylation status. The cell cycle inhibitory role of both p21 and p27 occurs

in the nucleus of the cell through interaction with CDKs, but many of their secondary functions

occur in the cytoplasm. For example the cyclin D-CDK assembly role occurs in the cytoplasm

(Sgambato et al., 2000).

Modulation of the p27 phosphorylation status can modify the protein binding domains leading to

changes in binding partners, and can alter the subcellular localization by revealing or hiding

nuclear localization and export sequences (Figure 1.12). Phosphorylation of p27 on Tyr

(74/88/89) alters the ability of p27 to interact and inhibit CDK2 whereas phosphorylation of p27

at Th187 by CDK2 provides a recognition motif for E3 ubiquitin ligases and targets p27 for

ubiquitination and proteosomal degradation, thus allowing further cyclinE-CDK2 complexes to

be activated (Besson et al., 2008). Subcellular localization of p27 is regulated by

phosphorylation at Ser10, Thr157 or Thr198. Phosphorylation at Ser10, the most common site of

p27 phosphorylation, reveals a binding site for CRM1/exportin, promoting p27 export from the

nucleus to the cytoplasm (Rodier et al., 2001;Connor et al., 2003;Besson et al., 2006). Further

phosphorylation of p27 at Thr157 or Thr198 site interferes with the nuclear localization sequence

effectively averting nuclear entry while concomitantly supporting its association with 14-3-3

which provides additional protein stability (Fujita et al., 2002;Sekimoto et al., 2004). P21 is

equally influenced by phosphorylation. Phosphorylation at various sites modulates the ability of

p21 to interact with cyclin-CDK complexes, PCNA, ubiquitination machinery and the

proteosome. In addition, phosphorylation of p21 can lead to its cytoplasmic translocation and

retention (Child and Mann, 2006).

1.3.2.2 Cell cycle regulators in cancer

Perturbation in p21 and p27 expression and function has been linked to several pathologies

associated with hyper-proliferation, including numerous cancers. Surprisingly however,

mutations of either p21 or p27 in human cancers are very rare (Slingerland and Pagano, 2000).

Knock out mouse models have verified that p21 and p27 have tumor suppressor roles with loss

of either inhibitor resulting in a predisposition to tumourogenesis. Surprisingly, overexpression

31

of cytoplasmic p21 or p27 is also a marker of poor prognosis for many cancers (Slingerland et

al., 2000;Besson et al., 2008). It has been suggested that increased p21 and p27 in the nucleus

may have tumor suppressive function by interacting with CDK, but that increased expression in

the cytoplasm independent of CDKs may be oncogenic (Besson et al., 2008). Additionally,

nuclear p21 has also been found to be tumor promoting by increasing the levels of cyclin D1 in

the nucleus (Liu et al., 2007).

Overexpression of D and E type cyclins also commonly occurs in many human cancers, where

expression is believed to accelerate G1 progression (Sherr, 1996). In contrast, cells lacking D

and E type cyclins are resistant to oncogenic transformation, as they have decreased ability to re-

enter the cell cycle from quiescence, and have difficulty responding to mitogenic stimulation

(Sherr et al., 2004;Kozar et al., 2004).

Although a great deal is known about general cell cycle regulation little is known regarding how

the cell cycle is regulated in the placenta: Which regulators are expressed and with which

functions are they associated? Which proteins interact, and how do these events change with

placental maturation? These questions are of great importance to our understanding of

placentation and even more important as a basis in which to compare placental pathologies in

which proliferation is altered.

1.3.3 Regulation of Proliferation and the Cell Cycle in the Placenta

Proliferation of the trophoblast during placental development is highly organized, being

restricted to the single layer of cytotrophoblast cells in the floating villi, and concentrated in the

proximal portion of the anchoring columns (Figure 1.4).

Progenitor cytotrophoblast cells in the floating villi proliferate, and give rise to daughter cells

that either aid in expansion of the chorionic villi during early placental development, or

differentiate and fuse with the outer syncytium to replenish the materials of the aging outer layer.

Exponential growth of the placenta early in the first trimester requires that a large percentage of

the cytotrophoblast population maintain their proliferative capacity. As gestation progresses the

rate of placental growth slows and the ratio of proliferative cytotrophoblast cells in the placenta

decreases, remaining predominantly to replenish the overlying syncytium of the floating villi and

allow for continual trophoblast turnover.

32

In the anchoring villi, proliferation of the trophoblast is critical to the establishment of extensive

cell columns that effectively anchor the placenta to the maternal wall. Cell cycle regulation in

this compartment is therefore extremely important, as an adequate number of EVT cells must be

produced and differentiate to become invasive, as to ensure sufficient spiral artery remodeling.

Interestingly, acquisition of an invasive phenotype is associated with a switch from a mitotic to

an endoreduplicative phenotype (Sherr, 1996;Zybina et al., 2004), a form of cell cycling that

entails continuous multiplication of the genome in the absence of mitotic division, resulting in a

polyploid cell. This process has been shown to be regulated in part by the E type cyclins in a

process that is CDK independent (Sherr, 1996;Geng et al., 2003;Parisi et al., 2003). Genomic

content in human EVT cells, found deep in the deciduas, can reach up to 8c-16c, and, in mice,

the DNA content of invasive trophoblast giant cells can reach up to 1000N (Barlow and

Sherman, 1972;Zybina et al., 2002). The occurrence of endoreduplication has been shown to be

an essential aspect of placentation, as defects in the ability of the trophoblast to endoreduplicate

in mice (a phenotype associated with cyclin E null mice) lead to placental associated embryonic

lethality (Geng et al., 2003;Parisi et al., 2003). Although the functional relevance of reaching a

polyploid state is not well understood, it has been proposed that endoreduplication acts as a

mechanism to allow sufficient gene transcription and assure that metastatic transformation is

prevented during normal pregnancy (Zybina and Zybina, 2005).

Abnormalities in cell proliferation in either the floating or anchoring villi can lead to improper

placental function and clinical aspects of disease. Excessive proliferation in the floating villi can

enhance the rate of trophoblast turnover and subsequently increase the amount of syncytial

material entering the maternal blood stream. Elevated levels of placental debris in the maternal

circulation have been shown to be a significant contributor to the pathogenesis of preeclampsia

(Levine et al., 2004). Likewise insufficient production of villous cytotrophoblast can lead to a

deficiency in syncytial renewal and impact upon the capacity of the syncytium to uptake the

necessary gas and nutrient requirements of the growing fetus, a phenomenon associated with

intrauterine growth restriction (IUGR).

In the anchoring villi the balance between cell proliferation and trophoblast differentiation is also

critical. Decreased proliferation, premature differentiation or arrest of the EVT in an immature

state may result in an insufficient number of the migratory and invasive trophoblast cell type,

resulting in shallow invasion of the endometrium and consequently relatively few trophoblast

33

cells reaching the spiral arteries. Poor remodeling of the spiral arteries prevents the influx of

oxygen needed for proper development, a feature common to a number of placental pathologies,

including preeclampsia and IUGR. On the other hand, if proliferation and differentiation of the

extra villous trophoblast occur in excess, an increased degree of invasiveness can prevail, a

critical factor in the pathogenesis of placental acretia, molar pregnancy and choriocarcinomas.

The trophoblast biology in placental pathology will be further discussed in section 1.4 and

section 1.5 of the Introduction.

Change in the proliferative status of the placenta is driven largely by the environmental

surroundings experienced by the placenta over gestation. In the first trimester, the low oxygen

environment promotes trophoblast proliferation (Genbacev et al., 1996;Burton, 2009).

Proliferation is also supported by growth factors, and cytokines including vascular endothelial

growth factor, epidermal growth factor, Activin, and TGF3 among others. Elevated oxygen

levels, such as that experienced by the placenta in the late first trimester, and transcription factors

such as glial cell missing 1 (GCM1) inhibit villous and extra villous trophoblast proliferation and

promote differentiation (Baczyk et al., 2009). Although the oxygen and growth factor milieu are

capable of promoting or inhibiting the proliferative capacity of the trophoblast, little is known

regarding the specific cell cycle machinery targeted by these factors.

Although a number of papers have reported on the expression of cell cycle regulators in the

placenta, the studies have been based primarily on immunohistochemical analysis and have

resulted in a number of conflicting findings. To date, it has been reported that cyclin D1, cyclin

D3, cyclin E1, p21 and p27 are expressed by the cells of the placenta, and that their expression

often changes between the 1st and 3

rd trimester. In contrast it was found that cyclin D2 was not

expressed by the trophoblast (DeLoia et al., 1997). The localization of the individual molecules

and the pattern by which their expression changes over development remains ill defined.

For example DeLoia 1997 reported that expression of cyclin D1 was exclusive to the villous core

and to the extravillous trophoblast and that its expression was increased with advancing

gestation; whereas others have reported that cyclin D1 expression is predominant in the

cytotrophoblast layer and that cyclin D1 expression decreases towards term (DeLoia et al.,

1997;Genbacev et al., 2000;De Falco et al., 2004). Similarly cyclin E1 was shown to be

expressed predominantly in the cytotrophoblast layer in two reports, but in the

34

syncytiotrophoblast layer by another study (DeLoia et al., 1997;Bamberger et al.,

1999;Genbacev et al., 2000). Conflicting studies regarding subcellular localization of protein

expression have also been reported. For example, cyclin D3 was observed as cytoplasmic in a

study by Genbacev et al and nuclear by De Loia et al. Since contradictory observations have

been made for p21 and p27 where their expression was reported in the either the CT and EVT or

in the ST (Bamberger et al., 1999;Genbacev et al., 2000;Korgun et al., 2006). These

contradictions may be resolved by increased samples sizes and the use of multiple analytical

methods in future studies.

To date the majority of studies examining the regulation of the cell cycle in the placenta have

been based on immunohistological derived data. Quantitative assessment of mRNA and protein

expression is necessary to better comprehend cell cycle regulation in the placenta. In addition it

would be of value to examine multiple molecules in the same study to gain a more complete

understanding of the dynamics of the G1 phase in placental development. The balance between

the cyclins and the CDK inhibitors is critical to the decision of cell fate. It would therefore be of

great importance to determine the relationship between the cyclins and inhibitors with respect to

their co-expression and interaction in the trophoblast cells. A more in depth analysis, linking

mRNA, relative protein expression, localization, protein interaction, and environmental

influence, in a multi-parametric analysis, discriminating between early and late first trimester

may clarify the current knowledge and provide a better understanding of how proliferation and

the G1 phase of the cell cycle are governed in the human placenta.

1.3.4 Role of Bcl-2 Family Members in Cell Fate

Recent evidence has revealed a functional role for several Bcl-2 family members in regulating

cell cycle progression. Overexpression of anti-apoptotic multi domain members, Bcl-2, BCL-xL,

and BCL-w, has been shown to inhibit passage through the cell cycle by arresting cells at G0/G1

(inhibiting cell cycle entry) and by delaying the progression to S phase (prolongs G0 or G1)

(Jamil et al., 2005;Zinkel et al., 2006). Furthermore, this effect has been associated with

increased levels of p27. The mechanism linking Bcl-2/Bcl-xL to p27 overexpression has not

been elucidated; however it has been hypothesized that it may be a result of caspase inhibition

(Zinkel et al., 2006). The pro-survival Mcl-1 is also anti-proliferative; however, unlike Bcl-2 or

Bcl-xL, its cell cycle function is manifested in the S and G2 phases. Mcl-1 can bind and inhibit

35

PCNA or enter the nucleus and bind CDK1 thereby preventing its activity (Fujise et al.,

2000;Jamil et al., 2005). On the other hand, Bax, a pro-apoptotic family member, appears to

confer an advanced rate of the cell cycle by mechanisms that are currently unknown (Brady et

al., 1996;Knudson et al., 2001). Interestingly, this has been associated with decreased levels of

p27 conferring increased CDK2 activity (Zinkel et al., 2006). Only one study has reported on

Mtd with respect to the cell cycle. Interestingly, this in vitro study found that the Mtd promoter

could be activated by the E2F1/3 transcription factor (Rodriguez et al., 2006), suggesting that

Mtd expression may be initiated at the G1/S boundary and have an effect on the cell cycle.

However, whether Mtd has a direct role in cell cycle regulation remains to be established and

warrants further studies. As depicted in Figure 1.13, anti-apoptotic molecules generally exhibit

anti-proliferative properties while pro-apoptotic molecules appear to promote cell cycle

progression (Figure 1.13). This would suggest that like the pro-apoptotic molecule Bax, Mtd

may function to advance the cell cycle forward.

Caspase involvement in the cell cycle is highly cell type specific, involving particular caspases

and specific target substrates (Lamkanfi et al., 2007;Timmer et al., 2007). The main caspases

involved in cell cycle regulation appear to be caspase-3, 8 and, to a lesser extent caspase 6, and

their target substrates include p21, p27, Wee and NF-AT. Caspase-8 has been found to promote

the proliferation of T cells through the cleavage of the CDK1 inhibitor Wee (Alam et al., 1999).

Caspase-8 is also involved in trophoblast cell differentiation in the placenta (Huppertz et al.,

2004;Launay et al., 2005). So far, caspase-3 has been found to promote cell cycle progression in

lymphoid cells, forebrain cells and keratinocytes. In lymphoid cells this has been attributed to

caspase cleavage of p27. In contrast, Caspase-3 has been found to inhibit proliferation of B cells

through a mechanism involving cleavage of p21 (Waga et al., 1994;Woo et al., 2003). No study

has investigated the cleavage of p21 and p27 by caspase-3 in the human placenta during

development.

1.4 Placental Pathology

A number of placental pathologies are associated with an altered balance between proliferation

and cell death of the trophoblast cells. These changes influence the capacity of the trophoblast to

undergo cell turnover or invasion, two important events in normal placental function. Thus,

research directed at examining the mechanisms governing trophoblast cell cycle and apoptosis

36

Ray et al., Placenta, 2008

Figure 1-13: Dual role of Bcl-2 family members in cell death and proliferation

Anti-apoptotic molecules Mcl-1, Bcl-2 and Bcl-xL prevent apoptosis and exhibit anti-proliferative properties while

pro-apoptotic molecules Mtd, Bax, and Bak facilitate apoptosis and appear to promote cell cycle progression. Figure

modified from Placenta, Ray et at, 2008.

37

will provide insight into the underlying mechanisms contributing to placental disease. Our

research is aimed at investigating the molecules involved in trophoblast cell fate, in the hope of

uncovering potential targets for prevention, detection, and treatment.

1.4.1 Preeclampsia

Preeclampsia (PE) is one of the most commonly faced complications of pregnancy seen by

physicians in the obstetric field, affecting 5-7% of all pregnancies. This disorder, unique to

human pregnancy, is also the leading cause of fetal and maternal morbidity and mortality

worldwide. Although a great deal of research has been conducted to comprehend its

pathophysiology, preeclampsia remains difficult to predict and diagnose at an early stage.

Additionally, few therapies are available with delivery often being the only option.

1.4.1.1 Clinical Detection and Classification of Preeclampsia

Preeclampsia is a heterogeneous, multi-faceted disorder that is characterized more by its

symptoms then its pathophysiology. Its diagnosis is based primarily on the sudden onset of

maternal hypertension (>160 mmHg systolic pressure or >110 mmHg diastolic pressure), and

proteinuria (dipstick of 3+ on two random urine samples collected at least 4 hours apart) in the

second half of pregnancy ( 2002). Cases may also present with persistent cerebral symptoms

(altered mental status, headaches, blurred vision, or blindness), swelling, edema, epigastric or

right-quadrant pain with nausea or vomiting, thrombocytopenia (platelet count of <100,000 l),

abnormal liver enzymes, or seizures (Sibai et al., 2005). Importantly, the current criteria can only

be used reliably to diagnose women after the 20th

week of gestation, when the disorder has

already manifested.

Since the disorder is heterogeneous, the extent or pathogenesis of preeclampsia may vary,

resulting in very different patient outcomes. Outcome is usually favorable for both mother and

fetus if the symptoms are mild and present after the 36th week of gestation (Hauth et al.,

2000;Sibai, 2003), while patients who develop symptoms prior to the 33rd

week are at higher risk

of fetal and maternal complications (Sibai, 2003;Habli et al., 2007), such as pre-term birth,

placental abruption, reduced amniotic fluid for the fetus, and fetal growth restriction.

Complications to the mother also occur, including liver failure, seizures, stroke and even death

(Sibai et al., 2005). Although the symptoms of preeclampsia dissipate with the completion of

38

pregnancy, the effects of the disease persist, and are associated with increased incidence of

cardiovascular problems in both the fetus and the mother in later life (Barker, 2003;Wilson et al.,

2003;Kajantie et al., 2009). Patients who experience hypertension with elevated liver enzymes

and low platelets are currently categorized in an independent category termed HELLP syndrome

(hemolysis with elevated liver enzymes and low platelets). This is an extreme form of

preeclampsia with increased risk to the mother and child. The extent of preeclampsia has been

theorized to reflect different underlying mechanisms potentiating the disease, possibly arising

from both fetal and/or maternal origins. Therefore, determining the molecular differences

between the classes of preeclampsia may aid in their differential diagnosis and treatment.

Of note, women who are obese, have preexisting hypertension or diabetes, have undergone

assisted reproduction, carry multiple fetuses, are above the age of 40 or who have had

preeclampsia in a previous pregnancy are at a higher risk of developing preeclampsia in the

current pregnancy. Many of these risk factors are on the rise in our society, foreshadowing a

potential increase in the incidence of preeclampsia in the future. Not only will this translate to an

increased number of woman and children that suffer, but it will put a financial strain on the

health care system. This amplifies the need for a better understanding of the biology of the

disease in order to better prevent, diagnose, and treat the syndrome.

1.4.1.2 The Preeclamptic Placenta and Trophoblast Biology: Cell proliferation

and Trophoblast Turnover

Although the etiology of preeclampsia remains unknown, evidence confirms that the placenta

plays a central role in its pathogenesis, as removal of the placenta remains the only effective

treatment for the disease. Furthermore, preeclampsia can arise in the absence of a fetus or uterus

(if placental material persists after pregnancy, or in the case of an ectopic pregnancy), and it is

more common in cases of increased placental volumes, such as in multiple gestation pregnancies.

However, even though the placenta is the key component in the pathogenesis of the disorder, it is

accepted that preeclampsia is a syndrome of vascular endothelial dysfunction and excessive

systemic inflammation.

Normal pregnancy is associated with shedding of syncytial particles into the maternal blood

stream as part of the natural syncytial renewal, and this is associated with a low state of

inflammation in the pregnant woman (Sargent et al., 2003;Aly et al., 2004). In preeclamptic

39

women the inflammatory response is exaggerated (Redman et al., 1999;Sargent et al., 2003). It is

believed that factors originating from the placenta irritate the maternal endothelium and cause

the clinical symptoms. Research has attempted to decipher the factor or factors released by the

placenta that stimulates the endothelial dysfunction and the heightened inflammatory response in

preeclampsia. Angiogenic factors including sFlt1 (Nevo et al., 2008) and endoglin, syncytial

microvillous membrane particles, trophoblast specific proteins (cytokeratin), and increased

concentration of fetal proteins and free fetal DNA (Zhong et al., 2001) have been suggested as

potential stimulators (Levine et al., 2004). The exaggerated inflammatory response seen in

preeclamptic women is believed to result from either the excessive release of placental factors

into the maternal circulation (placental origin), or from a sensitivity of the endothelial lining in

the maternal vasculature, decreasing its tolerance to the normal load of placental debris

(susceptible women). In cases of placental origin these factors may arise from an increased

placental deportation into the maternal circulation. This has been supported by studies that have

found increased placental debris in the maternal serum of preeclamptic patients (Johansen et al.,

1999;Ishihara et al., 2002;Sargent et al., 2003). Moreover, this material has been shown at 16-18

weeks, preceding the classical symptoms of preeclampsia (Levine et al., 2004). The underlying

reason for the increased placental debris seen in preeclamptic women however, is currently

unclear.

It is widely believed that the development of preeclampsia is initiated in the first trimester of

pregnancy when the process of trophoblast differentiation begins. It has been suggested that the

entire event of trophoblast turnover, from trophoblast proliferation through fusion to extrusion, is

increased in preeclampsia (Huppertz et al., 2004). However, while studies have investigated the

apoptotic component of this path, fewer studies have been done to determine if fusion or

cytotrophoblast proliferation is altered in preeclampsia. The few studies that have looked at the

proliferative status in preeclamptic placentae have shown, based on ki67 or BrdU staining, an

overall increase in cytotrophoblast proliferation (Arnholdt et al., 1991;Brown et al., 2005).

However, the mechanisms leading to the hyperplasia seen in preeclampsia have been overlooked.

Understanding these underlying events in trophoblast turnover could aid in distinguishing the

differences between placental and maternal origin of preeclampsia. In addition this line of

investigation also has potential to lead to novel diagnostic and therapeutic discovery.

40

1.4.1.3 The Preeclamptic Placenta and Trophoblast Biology: Spiral Artery

Remodeling and Oxygenation

Preeclamptic placentae are also associated with altered development of the extravillous

trophoblast cells and anchoring column formation. The EVT cells of preeclamptic cases are

characterized as immature compared to normal placentae. They express markers associated with

a proliferative phenotype and display decreased markers of trophoblast differentiation

(Arkwright et al., 1993;Redline and Patterson, 1995;Zhou et al., 1998;Caniggia et al., 1999).

Moreover, they remain hyperproliferative, with limited migration into superficial decidua

(Redline et al., 1995). The shallow trophoblast invasion in preeclampsia is also associated with

an over-expression of HIF-1, a regulator of hypoxia, and TGF3, an inhibitor of trophoblast

differentiation (Caniggia et al., 1999). Importantly, both the low oxygen environment and

elevated expression of TGF3 have been shown to promote the maintenance of the proliferative

phenotype (Genbacev et al., 1996;Caniggia et al., 1999;Caniggia et al., 2000). In addition,

preeclampsia has been associated with increased EVT cell death at the feto-maternal interface

preventing adequate cell invasion (Genbacev et al., 1999;DiFederico et al., 1999). These

phenomena lead to reduced decidual invasion, and as a result fewer maternal spiral arteries are

remodeled (Gerretsen et al., 1981;Zhou et al., 1997). Furthermore, those maternal arteries that

are remodeled do not reach as deep as those in normal pregnancy, and in many cases the smooth

muscle is not fully eroded. This subsequently results in the persistence of high-resistance

vasculature that remains sensitive to endocrine stimuli. It has been suggested that contractility of

the uterine arteries may persist causing the placenta to experience interval changes in oxygen or

ischemia reperfusion (a hypoxia reoxygenation) type event. Not surprisingly, hypoxia

reoxygenation and oxidative stress are among the major drivers of preeclampsia (Jaffe et al.,

1997;Hubel, 1999;Burton and Jauniaux, 2004). In addition, the narrow, high-resistance

vasculature, is often unable to deliver an adequate blood supply to the fetoplacental unit.

Consequently, placentae from preeclamptic women are often associated with oxygen and nutrient

deprivation to the placenta and fetus (Kingdom and Kaufmann, 1997). Furthermore, since low

oxygen is known to promote proliferation and reduced invasion in early placental development,

this environment may reinforce the maintenance of the hyperproliferative premature state.

Importantly, the high velocity perfusion, caused by insufficient spiral artery remodeling, has

been proposed as a mechanism leading to increased syncytial shedding in preeclampsia (Crocker,

2007).

41

Due to its complexity and heterogeneity, preeclampsia has presented as a challenging disorder to

diagnose and treat, with the current management consisting of supportive care and expedited

delivery. Even though a substantial amount of research has been done in the field, we are just

beginning to uncover some of the mechanisms underlying the disorder. To date, there have been

a number of reports focusing on the factors governing the invasive pathway of the trophoblast;

however, the mechanisms governing regulation of cell proliferation and cell cycle regulation in

the trophoblast have not been well established. It is of great importance that the molecular

mechanisms underlying the regulation of the cell cycle during normal and abnormal placentation

be determined. Understanding these events will allow for earlier detection and possible avenues

of treatment. Providing easier and more effective treatment strategies will alleviate the cost to

heath care, and most importantly assure a better wellness of life to the mothers and newborn

infants.

1.4.2 IUGR

Intra uterine growth restriction (IUGR) describes the failure of a fetus to reach its growth

potential. Importantly, birth weight is the strongest known indicator of perinatal morbidity and

mortality and, as such, IUGR presents as a serious complication of pregnancy (Pollack and

Divon, 1992). The pathology of IUGR is closely related to preeclampsia, sharing similar risk

factors and perinatal outcomes‟, however, the two are physiologically different disorders (Villar

et al., 2006). The distinction between the two conditions has recently become an important issue

in the obstetric field as it is crucial to the development of proper diagnostic and treatment

strategies.

1.4.2.1 IUGR classification and etiology

The complication in prediction, treatment and management of IUGR stems from the fact that the

disorder can arise from a number of factors. Simply, IUGR can result from fetal, placental, or

maternal causes. Fetal factors include genetic disorders, comprised mainly of chromosomal

abnormalities, as well as fetal infection and congenital malformations (Pollack et al., 1992).

Maternal contribution stems from factors that influence the ability of the mother to produce and

deliver an adequate source of oxygen and nutrition to the growing fetus. These include maternal

nutritional deprivation, disorders that are associated with hypoxia (eg: asthma, cyctic fibrosis,

lung or heart diseases), and vascular pathologies (preeclampsia, diabetes, chronic hypertenstion).

42

Environmental factors such as cigarette smoking and alcohol consumption also predispose the

fetus to intra uterine growth restriction as they impair oxygen and nutrient delivery (Pollack et

al., 1992).

Unexplained cases of IUGR (those not due to congenital malformation, structural defects,

maternal smoking or under-nutrition, or occur secondary to preeclampsia or gestational

hypertension), are often attributed to placental insufficiency (Villar et al., 2006). Importantly,

placental insufficiency is recognized as the most common cause of IUGR among healthy non-

smoking women with adequate nutritional status (Villar et al., 2006). These cases of IUGR arise

from defects that affect placental development, structure and function. In many cases IUGR is

associated with poor trophoblast invasion and remodeling of the maternal spiral arteries, which

affects the utero-placental circulation and subsequently gas and nutrient bioavailability. Reduced

oxygenation can be determined based on normal umbilical arterial blood flow by Doppler

analysis, where absent end diastolic flow and reverse end-diastolic flow are defined as abnormal,

and are associated with high fetal risk of IUGR (Ferrazzi et al., 2002). Poor placental

differentiation including decreased branching of the villous tree and altered composition of the

villous membranes are also seen in cases of IUGR. These defects can result in deficient placental

transport and impaired gas and nutrient up-take (Cetin et al., 2004;Cetin and Alvino, 2009).

Ultimately, these structural and functional changes result in insufficient nutrient supply and

absorption, resulting in reduced tissue deposition to the fetus (Cetin et al., 2004).

Although classification of IUGR is under debate (Pollack et al., 1992;Maulik, 2006), the most

commonly used criteria is based on a birth weight below the 10th

percentile when corrected for

gestational age and fetal sex, as well as Doppler analysis of the utero-placental circulation.

Additionally, babies whose birth weight is below the 5th

percentile are considered to be severely

IUGR.

The criteria for classification of IUGR have been controversial due to population diversity, and

variation in methods of gestational age and fetal growth measurement among institutions

(Pollack et al., 1992). This has resulted in the inadvertent incorporation of babies that are small

for gestational age (SGA) but relatively healthy, into the pathological arm of numerous studies.

Additionally, the time of onset of the disorder greatly impacts the outcome of the fetus and

should be taken into consideration within studies. IUGR at term may be apparent only as fetal

43

distress requiring a cesarean section and lead to a favorable neonatal outcome. In contrast, early-

onset IUGR before 33 weeks may be so severe as to cause fetal death in-utero, or require

immediate delivery to avoid fetal demise (Villar et al., 2006). These severe cases of IUGR are

often delivered preterm and are thus associated with secondary handicaps and require intensive

care. In addition, IUGR can also occur as a secondary disorder to preeclampsia. It has thus been

suggested that studies take precaution to differentiate between the individual groups. Recently,

an international study of 24,678 pregnancies compared the risk factors and perinatal outcomes

(fetal death, preterm delivery, and length of stay in ICU) of preeclampsia, gestational

hypertension and IUGR, and concluded that although preeclampsia and gestational hypertension

showed many similarities, unexplained (due to placental insufficiency) IUGR presented as a

different entity from preeclampsia (Villar et al., 2006). This paper suggests that the underlying

causes of preeclampsia, IUGR and IUGR secondary to preeclampsia are likely very different,

and stresses the importance of treating these groups as separate disorders. This highlights the

importance of classifying these pathologies of pregnancy separately in scientific studies.

Babies that develop IUGR have an increased risk of developing cardiovascular disease,

metabolic syndrome, and diabetes in adult life (Barker, 1998). In addition, neurological

development problems in both preterm and term IUGR infants is significantly increased (Pollack

et al., 1992). Due to the heterogeneity of the disorder, IUGR presents as a complex and difficult

disorder to handle clinically, with current management of IUGR being largely patient specific

(Cetin et al., 2004). It is therefore important that diagnostic tools be developed to advance the

efficiency and effectiveness of individualized care. Lastly, it has become increasing apparent that

the distinction between preeclampsia and IUGR is of great importance. Molecular

characterization of the two conditions will provide a better understanding of their biological

differences and provide insight into designing improved diagnostic, prevention and treatment

strategies.

1.4.3 Mtd and the Bcl-2 family in preeclampsia

The rate of apoptosis and trophoblast turnover is increased in preeclampsia relative to normal age

matched control placentae, indicating that molecules involved in these pathways may be

candidate markers of placental disease. Interestingly however, the expression of most Bcl-2

family members including Bcl-2, Bcl-xL, Bax and Bak in preeclampsia and preeclampsia

44

associated with IUGR remain unchanged compared to controls (Allaire et al., 2000;Levy et al.,

2002). In contrast, the expression of both Mtd and Mcl-1 are both altered in preeclampsia,

underscoring a unique relationship of these Bcl-2 family members in placental pathology. In

preeclampsia, the full length Mcl-1 protein is cleaved into a pro-apoptotic fragment and the

levels of both Mtd-L and Mtd-P are increased (Soleymanlou et al., 2005b;Soleymanlou et al.,

2007). Hence, in preeclampsia, the Mtd/Mcl-1 rheostat is tilted towards the production of pro-

apoptotic “killer” isoforms. One of our objectives, therefore, has been to decipher the role of Mtd

in this pathological condition. Recently we have discovered that in preeclamptic samples Mtd is

not only abundant in the apoptotic syncytial knots as we have previously shown, but that Mtd

expression also occurs in the proliferative cells of the cytotrophoblast layer (Soleymanlou et al.,

2005b;Ray et al., 2008). This may suggest that the increased Mtd/Bok expression seen in

preeclampsia may contribute to both the increased apoptosis and hyperproliferative nature of the

disorder.

Preeclampsia is associated with an increased expression of the p53 transcription factor (Heazell

et al., 2008b) which has been shown to be a transcriptional regulator of Mtd, and regulates

apoptosis and proliferation (Yakovlev et al., 2004). Preeclampsia is also associated with low

oxygenation and increased levels of Hypoxia Inducible Factor-1 (HIF-1) which has also been

shown to increase the expression of Mtd expression (Hubel, 1999;Hung et al., 2001;Soleymanlou

et al., 2005b).

1.5 Complete Molar Pregnancy

Complete molar pregnancy is a devastating condition whereby abnormal placental tissue

develops in the absence of a fetus. Not only is the pregnancy emotionally devastating, but it is

also associated with the development of further pathologies including preeclampsia and

choriocarcinoma. Although research has been conducted to comprehend the disease, the

pathophysiology of complete molar pregnancy remains unclear. Further research is required to

understand how molar pregnancy develops and how it persists, in order to better prevent and

treat the disease.

Molar pregnancy is not uncommon, occurring in approximately 1 in every 1500 pregnancies in

Europe and North America (Steigrad, 2003) and 1 in 200 in Asia (Seckl et al., 2000;Tham et al.,

2003). Although there are no known risk factors for developing molar pregnancy, the disorder

45

has been shown to be increased in women over 45, and in teenagers (Palmer, 1994;Paradinas et

al., 1996;Steigrad, 2003). Furthermore, developing a hydatidiform mole does put the woman at

an increased risk of developing a future molar pregnancy; 5-40 times more likely than a person

who has not developed the disease previously (Palmer, 1994;Paradinas et al., 1996;Steigrad,

2003). Most importantly, molar pregnancies are predisposed to malignant transformation with 8-

30% of patients developing gestational trophoblastic diseases (GTD) requiring chemotherapy

(Kurman, 1991a;Mazur and Kurman, 1994). Choriocarcinoma (metastatic cancer) one of the

most aggressive types of cancer affecting women, occurs in 2-3% of patients following a molar

pregnancy, an incidence 2000-4000 times greater than that following a normal pregnancy (Seckl

et al., 2000;Li et al., 2002). Due to the devastating condition faced by women with molar

pregnancy and the high risk of GTD following the disease, it is important that precise detection

and diagnosis of molar pregnancy be available.

1.5.1 Clinical Detection and Classification of Molar Pregnancy

Until recently complete hydatidiform mole was diagnosed around 16-18 weeks of gestation with

symptoms including vaginal bleeding, passage of grape-like structures, abnormal growth of the

uterus, abdominal pain, severe nausea and vomiting, and detection of high hCG levels (human

chorionic ganadotropin) (Matsui et al., 2003;Slim and Mehio, 2007). Currently, with the use of

high resolution ultrasonography complete molar pregnancies are now diagnosed in the first

trimester (after the eighth week of gestation) (Kim et al., 2006). Since complete molar pregnancy

has no fetal component, these „pregnancies‟ are clinically terminated upon identification.

Persistent trophoblastic disease can occur following removal of the placental tissue. These cases

are identified by persistent or rising levels of hCG following molar evacuation. To date there are

no reliable markers to indicate that a mole will become persistent.

Molar pathology is classified as either partial or complete moles, a distinction made in 1977.

Complete hydatidiform moles (CHM) are typically diploid with all 46 chromosomes paternally

derived (Slim et al., 2007). It is believed that the androgenic mole arises from fertilization of an

anuclear „empty‟ ovum by either two sperm (dispermy) producing a diploid mole of either 46XX

or 46XY karyotype, or more commonly from fertilization of an empty ovum by a single sperm

that undergoes division after egg penetration (monospermy), producing a diploid mole of

karyotype XX (Slim et al., 2007). Either event results in a conceptus of paternal genomic origin,

46

which lacks the genetic contribution regulated by maternal imprinting. Both monospermic and

dispermic moles appear to have similar malignant potential (Li et al., 2002). Interestingly, recent

reports estimate that only 80% of CHM are diploid androgenic (60% monospermic, 20%

dispermic) with the remaining 20% arising from biparental genomic contribution (Slim et al.,

2007). Biparental moles are believed to develop from a defect in the maternal locus in the region

of 19q13.4 (Panichkul et al., 2005) and are consequently improperly imprinted .

In contrast partial hydatidiform moles (PM) are typically of triploid karyotype (XXX, XXY, or

XYY) arising from dispermy of a normal ovum (diantric) (Szulman and Surti, 1984). This results

in the addition of a second set of paternal chromosomes (two chromosome sets of paternal origin

and a third set of maternal origin) increasing the total genetic material and effectively doubling

the amount of paternally imprinted genes. Unlike complete moles, partial moles give rise to both

embryonic and extraembryonic tissue, however the fetus does not fully develop in these cases.

Although the placental portion of a partial mole is similar to that of a complete mole, partial

moles tend to be less invasive, with the majority of invasive moles being diploid (CHM) (Wake

et al., 1984).

1.5.2 Trophoblast Biology of the Complete Molar Placenta: Morphological

characteristics and Histopathology

Typically the complete mole (of 16-18 weeks) has been characterized by hydropic degeneration

of the placental villi and the absence of fetal tissue (embryo, cord, and amniotic membranes).

The villi are enlarged and surrounded by areas of excessive proliferation, with inner cistern

formation and lack of vascular structure. With diagnosis now occurring in the first trimester (8-

12 weeks of gestation), recent papers have begun to report on a new set of morphological

identification landmarks (Sebire, 2010). For example, the comparison of molar tissue between 8-

12 and 16-18 weeks has identified the presence of a primitive vasculature that later disappears

(Kim et al., 2006). Importantly the study of early gestation moles has provided morphological

clues regarding the development of the disorder, and suggests that further evaluation of molar

development spanning into late gestation would provide an even deeper insight in to its

pathogenesis. However, few second trimester specimens exist to be studied, since molar

pregnancies are artificially stopped once identified. Nevertheless, in rare molar cases placental

development is continued, if the mole develops in conjunction with a twin pregnancy. In these

rare cases, the molar placenta develops independently alongside a genetically normal twin

47

conceptus. These pregnancies are extremely high risk, and only with extreme caution are they

maintained to the earliest point of safe delivery.

Although new insights are being made through the study of molar tissue morphology, molecular

studies are required to gain a deeper understanding of the underlying mechanisms regulating this

type of placental development.

1.5.3 Trophoblast Biology of the Complete Molar Placenta: Molecular

Characteristics

Placentae from molar pregnancies exhibit excessive trophoblast proliferation and apoptosis. In

addition, Immunohistochemical studies have reported on abnormalities regarding the expression

of a variety of cell cycle regulating molecules in molar pregnancy. Although primarily

observational, studies agree that the highly proliferative phenotype of molar placentae is

associated with increased Ki67 and cyclin E expression (Kim et al., 2000;Olvera et al.,

2001;Bamberger et al., 2003). Furthermore, an increase in CDK2 and E2F-1 is also seen in molar

pregnancy (Olvera et al., 2001). p27 has been shown to be expressed by the molar placenta;

however, the literature is conflicting in regards to how it compares to normal placenta of the

same gestation (Olvera et al., 2001;Fukunaga, 2004). Surprisingly, p21 is upregulated in molar

pregnancy despite the hyperproliferative nature of the disorder (Cheung et al., 1998).

The study of proliferation markers in the mole has been shown to be helpful in differentiating

complete vs partial moles and distinguishing moles from spontaneous miscarriage (Xue et al.,

2005). However, research to date has suggested, that the proliferative indices of the mole have no

predictive value with respect to malignant transformation (Cheung et al., 1998). In contrast,

studies indicate that the incidence of apoptosis negatively correlates with progression to GTD

(Wong et al., 1999). In addition, it has also been theorized that increased stromal apoptosis may

be associated with inadequate vascularization contributing to the improper development of the

villi (Kim et al., 2006). It is therefore of great importance to determine the molecules involved in

regulating apoptosis in the molar placenta.

Surprisingly, very few reports have focused on apoptosis in molar pregnancy. Apoptosis,

typically detected by TUNEL recognition, has been found to occur at higher levels in complete

hydatidiform mole compared to partial mole, spontaneous abortion or normal pregnancy (Qiao et

48

al., 1998;Halperin et al., 2000;Chiu et al., 2001;Kim et al., 2006). Interestingly molar pregnancy

has also been found to be associated with increased levels of p53, a phospho-protein that

regulates both apoptosis and proliferation (Qiao et al., 1998;Halperin et al., 2000;Chiu et al.,

2001). Importantly, p53 regulates the expression of many Bcl-2 family members. This includes

the down-regulation of Bcl-2, and the up-regulation of both Bax, and Mtd (Miyashita et al.,

1994;Yakovlev et al., 2004). Although p53 is increased in the molar pathology, no significant

difference in Bax expression and only slightly lower Bcl-2 levels have been detected (Qiao et al.,

1998;Wong et al., 1999;Chiu et al., 2001). Furthermore, neither expression of Bcl-2 nor Bax has

been associated with progression to GTD. Additionally, no significant differences have been

detected at the protein level for apoptosis activating caspase 8 or caspase 10, however caspase 3

activity was shown to be increased in the stromal region of molar tissue (Chiu et al., 2001;Fong

et al., 2006;Kim et al., 2006).

Surprisingly, no study to date has investigated the expression of Mtd in molar pregnancy. Its

exploration however seems intuitive, given that Mtd is highly expressed in the placenta and that

molar pregnancy is associated with a high incidence of developing preeclampsia (Soleymanlou et

al., 2005b). Furthermore, molar pregnancy exhibits increased expression of HIF-1, and E2F-1,

two transcription factors that are hypothesized to induce Mtd expression.

Although cell cycle and apoptotic regulating molecules have been reported in the hydatidiform

mole, these studies have been primarily of descriptive nature and clinically focused. In addition

they have been predominantly restricted to molar tissue of the first trimester. In the future

additional molecular based studies on the mechanisms driving molar progression are needed, to

provide useful knowledge towards the development of diagnostic and treatment strategies.

Importantly the study of cell cycle progression and cell death in the molar placenta will provide

insight into the overall understanding of cell cycle regulation during trophoblast proliferation and

the mechanisms that determine trophoblast cell fate.

1.6 Thesis Hypothesis and Objectives

In humans, cellular proliferation, differentiation and death accompany early placental

development of the trophoblast lineage, the cells forming the placenta. Abnormality at any stage

49

of this development, due to altered proliferation, differentiation or cell death may lead to

improper placental function and subsequent pregnancy related complications.

Members of the Bcl-2 family are central to nearly all pathways governing apoptotic cell death.

Additionally, an increasing body of evidence suggests that a number of anti-apoptotic Bcl-2

members are also involved in the regulation of the cell cycle. Mtd, a pro-apoptotic member of the

Bcl-2 family, is expressed primarily in the placenta and is increased in pregnancies complicated

by preeclampsia. In addition it has been reported that the Mtd promoter is activated at the G1/S

boundary by the E2F1/3 transcription factor, indirectly implicating a role for Mtd during the

early stages of cell cycle progression.

Placental pathologies including preeclampsia, intra uterine growth restriction, and molar

pregnancy are characterized by an immature proliferative trophoblast phenotype accompanied by

excessive cell death. We have previously found Mtd to be associated with the increase in

trophoblast cell death in preeclamptic placentae, however, the relationship between Mtd

expression and the hyperproliferative nature of preeclampsia has not yet been studied.

Furthermore, the molecular mechanisms regulating cell cycle progression and cell fate in normal

and pathological placentation remains unclear.

My overall objective was to investigate the role of Mtd in placentation, focusing specifically on

its role in the G1 phase of the cell cycle. Secondly I aimed to distinguish the factors involved in

promoting the cell cycle in normal placental development and to establish the underlying

molecular defects associated with trophoblast proliferation in placental pathology.

My overall hypothesis was that the expression of Matador is associated with cell proliferation

and apoptosis in normal placental development and in pathological conditions, and that altered

cell cycle regulation at the G1 phase leads to hyperproliferation of the trophoblast and

contributes to placental pathologies including preeclampsia, IUGR, and molar pregnancy

The first objective, presented in chapter 2 of the thesis, was to establish whether Mtd was

directly involved in G1/S phase transition in the trophoblast and to investigate the mechanisms

by which Mtd differentially regulated proliferation and apoptosis. This chapter uncovered a

direct effect of Mtd-L on cyclin E1 expression in proliferative trophoblast cells in normal

placental development.

50

This led to the second objective (Chapter 4) which examined how the G1 phase CDK activator,

cyclin E1 and the CDK inhibitor, p27 were regulated in normal trophoblast development and in

preeclampsia.

Finally, the fifth chapter examined the expression of Mtd in molar pregnancy. Conditions of

molar pregnancy have a high incidence of developing preeclampsia, and exhibit increased

expression of Hif-1 and E2F-1, two transcription factors that are hypothesized to induce Mtd

expression. Therefore, the final objective presented in this thesis (chapter 5) was to determine

whether molar tissue displayed increased levels of Mtd associated with apoptosis and altered

regulation of molecules involved in the G1 phase of the cell cycle.

51

2 Materials and Methods

2.1 Placental Tissue Collection

Informed consent was obtained from each individual patient. Tissue collections were approved

by the Mount Sinai Hospital's Review Committee on the Use of Human Subjects and carried out

in accordance with the participating institutions' ethics guidelines, and in accordance with the

guidelines in The Declaration of Helsinki. First-trimester human placental tissue (5-7 weeks of

gestation, n=30; 9-14 weeks of gestation, n=35) were obtained immediately following elective

termination of pregnancies by dilatation and curettage, or suction evacuation.

2.1.1 Placental samples for studies on preeclamptic pathology

Preeclamptic placentae were selected to represent classic severe early onset preeclampsia (PE;

n=47), late onset preeclampsia (LPE; n=20), and severe preterm Intra Uterine Growth Restricted

(IUGR; n=28) cases, according to the clinical and pathological criteria set out by the American

College of Obstetrics and Gynecology (ACOG 2002; Abuhamad, 2008). Placentae from IUGR

pregnancies were documented with absence or reversal of end diastolic velocity in the umbilical

artery without signs of preeclampsia, and were of fetal weight less than 5th

percentile. Preterm

normotensive age-matched control placentae (AMC; n=52) were selected as age-matched healthy

pregnancies with normally grown fetuses that did not have signs of placental dysfunction. Term

samples (TC; n=51) were obtained immediately following delivery and were within healthy

physiologic range for all maternal and fetal clinical parameters. Birth weight, gestational age,

laboratory values and clinical observations relevant to the health of the mother were taken from

the clinical records. Patients with diabetes, infections and kidney disease were excluded. Due to

organ heterogeneity, multiple specimens were sampled from central and peripheral regions of the

placentae. Placenta samples with calcification, necrosis and visually ischemic areas were also

excluded from the collection. Normal preterm deliveries were due to either multiple pregnancy

without discordancy, preterm labour due to incompetent cervix and premature preterm rupture of

membrane. Term controls were included as pregnancies delivered preterm are by definition

abnormal and may be associated with placental abnormality (Hansen et al., 2000;Redline, 2008).

Tissues were snap frozen or prepared in paraformaldehyde (PFA) and processed for

histochemical analysis. Clinical parameters for severe early preeclamptic, late preeclamptic,

IUGR, preterm age-matched, and term placentae are summarized in Table 2.1.

52

Table 2-1: Clinical data for preeclamptic, intra uterine growth restricted, and control cases

Values are expressed as averages +/- SEM. (A.G.A: appropriate for gestational age, IUGR

intrauterine growth restricted. CS: caesarian section, VD: vaginal delivery)

53

2.1.2 Samples for studies on molar twin pathology

The study presented in the Chapter 5 includes two molar twin sets consisting of a complete

hydatidiform mole as well as a co-existing genetically normal placenta with fetus. Clinical

parameters for the molar cases are summarized in Table 2.2.

The group of control twin sets included placentae from normal twins without discordancy and

without signs of growth restriction in either of the twins. The twin sets were selected as age-

matched healthy pregnancies with normally grown fetuses that did not have signs of placental

dysfunction or preeclampsia. None of the twin pregnancies had evidence of twin to twin

transfusion syndrome (TTTS) and none of the control twins had evidence of infection, anomalies

or abnormal chromosomes. Chorionicity was determined by first trimester ultrasound scan and

confirmed after delivery by histopathological examination of the placenta and membranes. The

placentas of the monochorionic twins were sampled from the area of the cord insertion of each

twin and lateral to it, avoiding the area between the cord insertions in order to reduce the chances

of sampling from shared cotyledons. The clinical characteristics of the twin pregnancies are

shown in Table 2.2. Birth weight, gestational age, laboratory values and clinical observations

relevant to the health of the mother were taken from the clinical records. Placenta samples with

calcification, necrosis and visually ischemic areas were also excluded from the collection.

Tissues were snap frozen or prepared in PFA and processed for histochemical analysis.

2.1.3 Samples for laser capture microdissection

Samples for laser capture microdissection were prepared immediately after collection. Fresh

tissue was washed in phosphate-buffered saline (PBS), saturated in cryoprotect solution (66%

OCT in 33% sucrose solution) and embedded in OCT (optimal cutting temperature) compound

(Tissue-Tek®, Sakura Finetek, Torrance, CA), snap frozen and kept at -800C until further use.

Cryosections, (7µm thick) were cut using RNAse-free blades and mounted on to uncoated,

uncharged slides (Superfrost, Fisher Scientific, Ottawa, Ontario).

54

Table 2-2: Clinical data for molar twin pregnancies and age matched control twin cases

Values are expressed as averages +/- SEM. (CS: caesarian section, VD: vaginal delivery)

55

2.2 First trimester Villous Explant Culture

First trimester placental tissue were collected following elective termination, rinsed and stored in

ice-cold PBS, and processed within 2 hours of collection. Endometrial tissue and fetal

membranes were dissected out. Small fragments of placental villi (25-40 mg wet weight) were

carefully dissected, and placed in either serum-free DMEM/F12 media (GIBCO BRL, Grand

Island, New York, USA) supplemented with 100 g/mL streptomycin, 100 U/mL penicillin, or

placed on Millicel-CM culture dish inserts (Millipore, Bedford, MA) previously coated

with 150

µl of undiluted Matrigel (Collaborative Biomedical Products, Bedford, MA). Explants were then

incubated at 37oC in standard condition (5% CO2 in 95% air, 20% O2, 150 mmHg) overnight.

Treatment was performed the following day.

2.2.1 Mtd antisense knockdown

Chorionic villous explant culture and Mtd knock-down was performed using phosphorothiolated

(all positions) sense (S) and antisense (AS) oligonucleotides designed against the Mtd-L

transcript NM_032515 (S-L: 50-CATGGAGGTGCTGCGG-30, AS-L: 50-

CCGCAGCACCTCCATG-30) as previously described (Caniggia et al., 2000;Soleymanlou et

al., 2005b). Villous explants were incubated at 3% O2 (92% N2 and 5% CO2, 21 mmHg) in the

presence or absence of S or AS oligos (10 mM) or DMEM/F12 alone for 72 hours. Explants

from eight different first trimester placentae (6-8weeks) run in triplicate, were used for the

antisense knockdown experiments.

2.2.2 TGF treatment

To study the effect of TGF on trophoblast cell cycle, cultured explants were treated with either

10ng/ml of TGF1 or TGF3 (R&D Systems) or left untreated. Villous explants were then either

maintained at 37oC in either standard tissue culture conditions (5% CO2 in 95% air, 20% O2,

150 mmHg) or in an atmosphere of 3% (92% N2 and 5% CO2, 21 mmHg) or 8% (87%

N2 and

5% CO2, 57 mmHg) O2 for 24 h at 37°C. Explants from a single placenta were prepared in

triplicate for each treatment condition. Experiments from 3 different first trimester placentae (6-

8weeks) were conducted for analysis in each study.

56

2.3 Laser Capture Microdissection

Prior to laser capture microdissection, immunostaining for Ki67 was performed initially to

identify areas of proliferation. Frozen sections were fixed in 4% PFA, treated with 0.3%

hydrogen peroxide (H2O2), blocked with 5% horse serum, and incubated with anti-Ki67 antibody

(dilution 1:100) for 1 hour at room temp. Laser capture microdissection was performed on

adjacent cryosections, using the Arcturus Pixcell II system (Arcturus Engineering, Mountain

View, CA) according to the manufacturer‟s protocol. Sections were overlaid with a thermoplastic

membrane (CapSure LCM Caps, Arcturus, Mount view, CA), and cells were captured by focal

melting of the membrane through laser activation. Three individual caps were used for each

placental sample, specifically isolating either, the villous trophoblast layer comprising both

cytotrophoblast and syncytiotrophoblast cells; the proliferative (Ki67 positive) region of

proximal column; or the non-proliferative (Ki67 negative) cells of the distal anchoring columns.

Approximately 2000-5000 cells were captured onto each cap.

2.4 RNA Analysis

RNA was extracted from frozen placental tissue or collected cells, using either an Rneasy Mini

Kit (Qiagen, Valencia, CA) or a Trizol extraction method. RNA from LCM samples was

extracted using PicoPure™ RNA Isolation Kit (Arcturus). All samples were treated with DNase I

to remove genomic DNA. 1 g of purified RNA was then used to synthesize cDNA using

random hexamers (Applied Biosystems, Foster City, CA), (denature 5 minutes at 65oC;

amplification 10min at 25oC, 120 minutes at 37

oC, and five minutes at 95°C). Analysis was done

using the DNA Engine Opticon®2 System (MJ Research, Waltham, MA). Mtd was quantified

using the SYBR Green I dye DyNamoTM

HS kit (MJ Research) based on the manufacturer‟s

protocol using isoform specific primers (Mtd-L: Forward 5‟-GCCTGGCTGAGGTGTGC-3‟,

Mtd-P: Forward 5‟-GCGGGAGAGGCGATGA, Reverse (both L and P) 5‟-

TGCAGAGAAGATGTGGCCA-3‟). Taqman Universal MasterMix and specific Taqman

primers and probe for cyclin E1, p27, cyclin D1, cyclin D3, and 18S were purchased from ABI

as Assays-on-DemandTM

for human genes (Applied Biosystems, Foster City, CA). Data were

normalized against expression of 18S ribosomal RNA using the well established 2-CT

formula

as previously described (Livak and Schmittgen, 2001).

57

2.5 Antibodies

Rabbit polyclonal antibody generated against a peptide mapping within an internal region of Mtd

(137-151) of human origin (NM_032515) was raised in the Caniggia laboratory. Rabbit serum

collected prior to Mtd peptide inoculation (pre-immune serum) was used as negative control, to

identifiy non-specific bands by Western blotting (WB 1:500). Remaining antibodies were

purchased from Cell Signaling Technology, Beverly, MA: rabbit polyclonal Mtd/Bok (4521) [IF

1:50]; rabbit monoclonal Cytochrome c (136F3) [WB 1:200]; mouse monoclonal cyclin D1

(DCS6) #2926 [IF 1:300(pc), WB 1:1000], mouse monoclonal cyclin D3 (DCS22) #2936 [IF

1:300, WB 1:1000], and rabbit monoclonal p21 (12D1) #2947 [IF 1:200, WB 1:1000]. The

following antibodies were purchased from Santa Cruz Biotechnology, Santa Cruz, CA: rabbit

polyclonal Mtd/Bok (H151) [IF 1:400], mouse monoclonal Cyclin E1 (HE12) sc-247 [IF 1:400,

WB 1:1500], goat polyclonal Lamin A (C20) sc-6214 [WB 1:200], mouse monoclonal GFP (B-

2) sc-9996 [WB 1:500], rabbit polyclonal p27 (N-20) sc-527 [IF 1:300, WB 1:500], rabbit

polyclonal phospho-p27 (Ser10)-R sc-12939-R [IF 1:400, WB 1:200]; rabbit polyclonal

phospho-p27 (Thr-198) sc-130603 [IF 1:400, WB 1:200]; rabbit polyclonal CDK2 (M2) sc-163

[WB 1:200]; and goat polyclonal Actin (I-19) sc-1616 [WB 1:15,000]. Multiple bands were

observed by Western blotting for the p27 (N-20) antibody. These bands were verified by

preabsorption with a p27 blocking peptide (N-20P at 5X the dilution of the p27 antibody).

Antibodies were also purchased from Vector Laboratories, Burlingame, CA: mouse monoclonal

Ki67 (clone MM1) [IF 1:100]; and Sigma-Aldrich, St.Louis, MO: mouse monoclonal -Tubulin

(clone DM 1A) [WB 1:2000]; DakoCytomation, Denmark A/S: mouse monoclonal proliferating

cell nuclear antigen (PCNA; clone PC10) [WB 1:500]; Abcam, Cambridge, MA: mouse

monoclonal E-Cadherin (HECD-1) [IF 1:200]; Roche, Mannheim, Germany: mouse monoclonal

BrdU [IF 1:100]; and R & D systems, rabbit polyclonal phosphor-p27 (Th157) [WB 1:2000].

Normal rabbit and mouse IgG (sc-2027 and sc-2025 respectively) were purchased from Santa

Cruz Biotechnology and used as negative control. Secondary antibodies were purchased from

Santa Cruz Biotechnology: horseradish peroxidase-conjugated or biotinylated anti-

rabbit/mouse/goat IgG [WB 1:5000]; or Molecular Probes, Eugene, OR: Alexa Fluor 488 anti-

goat/rabbit, Alexa Fluor 594 anti-mouse/rabbit [IF 1:200].

58

2.6 Western Blot Analysis

Western blot analysis was performed on 30g of total protein separated on 12%, or 14% (wt/vol)

SDS-PAGE gels and transferred to PVDF membranes. Non-specific binding was blocked by a 60

minute incubation in 5% (wt/vol) nonfat dry milk in Tris-buffered saline containing 0.1%

(vol/vol) Tween-20 (TBST). Following manufacturer‟s protocol the membranes were incubated

overnight at 4oC with primary antibody diluted in either 5% (wt/vol) nonfat dry milk or 5%

(wt/vol) bovine serum albumin (BSA). Membranes were then incubated with secondary

antibodies (horseradish peroxidase-labeled) prepared in 5% (wt/vol) nonfat dry milk for 60

minutes at room temperature and visualized by enhanced chemiluminescence (Western

LightningTM

Chemiluminescence Reagent Plus, Perkin Elmer, Shelton CT, USA) exposure to x-

ray film (Kodak). All western blots were confirmed for equal loading using 0.1% (w/v) Ponceau

S solution. For quantification purposes, bands of interest were scanned using CanoScanLiDE90

image scanner (Canon Canada Inc. Mississauga, ON) and analyzed using Image Quant software

5.0 (Molecular Dynamics).

2.7 Immuno-precipitation

For interaction studies, 300g of total protein was incubated with either 2g of p27 antibody or

normal rabbit IgG at 4oC overnight. 30l of protein A agarose beads sc-2001 (Santa cruz, Santa

cruz, CA) were then added to each sample for 2 hours at 4oC. The samples were then spun down

and the pelleted beads subjected to a series of washes in RIPA, and 1 X PBS and a final 50l

volume of 2X SDS sample loading buffer (125mM Tris-HCL (pH6.8), 4% (w/v) SDS, 20%

glycerol, 0.005% (w/v) bromophenol blue, 10% -mercaptoethanol), and heated for 5 minutes at

95oC. 25l of the resulting supernatant was subjected to SDS-PAGE followed by western

blotting.

2.8 Peroxidase Based Immunohistochemistry

Placental tissue were rinsed in PBS, dehydrated in 70% Ethanol and fixed for 2-4 hours at 4oC in

4% (vol/vol) paraformaldehyde and embedded in paraffin. Sections of 7m width were cut.

Every 10th

section was stained with hematoxylin and eosin in order to verify the quality of the

tissue and to aid in the selection of the most representative sections. Sections were de-

paraffinized in 3 five minute immersions in xylene, re-hydrated in a serial gradient of alcohol

59

solutions (two minute immersions in 100, 100, 100, 95, 90, 85, 80, 75, 70, and 50% ethanol),

followed by a five minute immersion in double distilled water (ddH2O), and washed in

phosphate buffered saline (PBS). Sodium Citrate antigen retrieval was performed using 10mM

sodium citrate pH6.0 (Cyclin E1: slides in sodium citrate solution were subjected to 5 minutes at

power 4 in a microwave and left to incubate for 15 minutes, this was followed by a an additional

3 minutes at power 4 in the microwave and a 20-30 minute cooling period; Ki67; slides were

boiled in sodium citrate by pressure cooker for 10 minutes) and endogenous peroxidase enzyme

activity was quenched with 3% (vol/vol) hydrogen peroxide in methanol for 30 minutes. Non-

specific binding sites were blocked using 5% (vol/vol) normal horse serum (NHS) and 1%

(wt/vol) BSA in Tris-buffer for 60 minutes at room temperature. Slides were incubated overnight

at 4oC with primary antibody. The following day the slides were incubated with biotinylated

secondary antibody (1:200) for 60 minutes at room temperature and subsequently incubated with

avidin-biotin complex for 60 minutes. Following a final wash slides were subjected to 0.075%

(wt/vol) 3,3-diaminobenzidine tetraaminobiphenyl (DAB) in PBS (pH 7.6) containing 0.002%

(vol/vol) H2O2, initiating a reaction that produced a brownish product. The reaction was stopped

in PBS and the slides were then counterstained with hematoxylin. Prior to mounting the slides

were dehydrated in an ascending ethanol series, and cleared in xylene. In control experiments,

primary antibodies were replaced with non-immune IgG.

2.9 Immunofluorescence (IF) Staining

Sections were de-paraffinized in 3 five minute immersions in xylene, re-hydrated in a serial

gradient of alcohol solutions (two minute immersions in 100, 100, 100, 95, 90, 85, 80, 75, 70,

and 50% ethanol), followed by a five minute immersion in ddH2O, and washed in PBS. Antigen

retrieval was performed using 10mM sodium citrate pH6.0 (in sodium citrate solution slides

were subjected to 5 minutes at power 4 in a microwave and left to incubate for 15 minutes, this

was followed by a an additional 3 minutes at power 4 in the microwave and a 20-30 minute

cooling period). Slides were then washed in PBS followed by a 10min incubation with Sudan

Black (0.1% sudan black in 70% EtOH) to quench endogenous fluorescence typical of red blood

cells. After additional PBS washes sections were pre-incubated in 5% horse serum diluted in

antibody diluent (0.04% sodium azide and 0.008% gelatin in PBS), for 60 minutes at room

temperature to block non-specific binding. Slides were then incubated with primary antibodies

diluted in antibody diluent overnight at 4ºC. The following day sections were washed in PBS and

60

incubated with fluorescence conjugated 2o antibodies [Alexa Fluor®488 (donkey anti-mouse, or

donkey anti-rabbit), Alexa Fluor®594 (donkey anti-mouse, or donkey anti-rabbit)] diluted in

antibody diluent for 60 minutes at room temperature in a covered container. Slides were then

washed in PBS, submerged for ten minutes in 0.4% DAPI (4‟.6-diamidino-2-phenylindole) for

nuclear detection and mounted in 50% glycerol solution. For dual labeled slides, samples were

incubated simultaneously with two 1o antibodies raised in different species. 2

o antibodies were

conjugated to fluoroforms of different wavelengths (specified above) and applied separately for

60 minutes each. For negative controls, primary antibody was replaced by corresponding

concentration of mouse or rabbit non-immune IgG. Fluorescence images were viewed using 20x

regular and 40x and 100x oil immersion objective lens (NA 1.35) and collected using

DeltaVision Deconvolution microscope (Applied Precision, LLC, Issaquah, WA).

Positive and negative cell counts were based on the presence or absence of immunoreactivity of

the Ki67 and Mtd antibody. Cell counts were recorded as a percentage of the total cell number in

the field where the total cell number was taken as the number of nuclei (DAPI stained) in the

trophoblast layer of floating villi or within anchoring columns. For Ki67 and Cyclin E1

quantification in normal and pathological samples, placental sections were stained using a

peroxidase-based method and cell counts were recorded as a percentage of either Ki67 or cyclin

E1 positive trophoblast cells per field. Five fields of view were analyzed per sample.

2.10 TUNEL (Terminal Deoxynucleotidyl Transferase-dUTP-Nick End Labeling)

Paraffin sections were dewaxed in xylene, rehydrated in descending grades of ethanol and rinsed

in PBS. Tissue sections were then pre-treated with 10g/ml proteinase K (invitrogen) in PBS for

10 minutes and treated with 3% H2O2 in methanol to quench endogenous peroxidase activity.

Following PBS washes, sections were incubated with TdT solution (Amersham, Piscataway NJ)

[1X One-Phor-All buffer (100mM Tris-acetate, 100uM magnesium acetate, 500mM potassium

acetate), biotin-16-dUTP (Fermentas), 1M dATP (Fermentas), and 210units/ml of TdT

(terminal deoxynucleotide transferase) enzyme in 0.1% triton X-100 in ddH2O) at 37oC for 1.5

hours. For peroxidase based staining sections were then subjected to ABC reagent/solution

(Vector laboratories) followed by multiples washes in 0.1% Triton-X in PBS, and finally DAB

substrate (Vector laboratories). Sections were counterstained with Harris hematoxylin (Sigma-

61

Alderich) and subjected to acid ethanol (1% HCL in 70% Ethanol) and sodium bicarbonate (1%

in H2O) treatment prior to mounting. Images were captured with a LEICA DC 200 imager. For

dual fluorescence labeled staining, 1o antibody against Mtd was applied following TdT

enzymatic reaction and slides were left overnight at 4oC. Secondary antibody application and

mounting were performed as described for IF staining.

2.11 Cell Line Culture and Analysis

Human choriocarcinoma JEG-3 cells (ATCC, Manassas VA) were grown in EMEM media

(ATCC, Manassas VA) supplemented with 10% (vol/vol) fetal bovine serum (non-heat

inactivated) in standard conditions, 20% oxygen (5% CO2 in 95% air).

2.11.1 SNP (Sodium nitroprusside) treatment

JEG-3 cells were seeded into a 6 well plate (2 X 105 cells per well) or 96 well plate (5000

cells/well) and treated after 16 hours with SNP (Sodium nitroprusside, Sigma) (dose range

between 1mM and 10mM).

2.11.2 Trypan Blue Exclusion Assay

Cells were harvested by trypsinization, centrifuged, resuspended in media, and diluted 1:10 in

0.4% Trypan blue solution (Invitrogen Corp., Carlsbad, CA, USA). Blue vs. white cells were

counted on a hematocytometer. Data is represented as an average of three independent

experiments.

2.11.3 Cell Viability (MTT)

Cell viability was determined by the 3-[4,5-dimethyl-thiazol-2-yl]-2,5-diphenyltetrazolium

bromide (MTT, Sigma) dye-reduction assay. JEG-3 cells seeded into a 96 well plate (5000 cells

per well) were subjected to 30ul of MTT dye (5mg/ml) per well, at 0, 6, 12, 24, and 48 hours.

The cells were then incubated at 370C for 4 hours. The media was subsequently aspirated and

100ul of dimethyl-sulfoxide (DMSO) were added to each well. Absorbance (relative optical

density) was measured at 570nM with an uQuant microplate spectrophotometer (Bio-Tek

Instruments, Winooski, VT). Data is presented as an average of four experiments.

62

2.11.4 Cell Fractionation

JEG-3 cells were grown to 70% confluence and collected in lysis buffer (0.3M sucrose, 1M

EDTA pH8, 5mM Mops, 5mM KH2PO4, 0.1%BSA – pH to 7.4 using KOH) for fractionation.

Cells were lysed using a dounce homogenizer, and differentially centrifuged to attain cell

fractions: 5 min at 600g to precipitate nuclei followed by 12 min at 10,000 g to pellet light

membrane fraction. Resulting supernatant containing the cytoplasmic fraction was lyophilized

and all fractions were resuspended in RIPA buffer. Verification of fraction specificity was

assessed by Western blot detection of specific subcellular markers: -tubulin (cytoplasm),

cytochrome c (mitochondria), and lamin A (nuclear membrane).

2.11.5 Localization of Mtd to Mitochondria

3.0x105 JEG-3 cells were seeded on sterile glass cover slips and allowed to adhere overnight.

Cells were then incubated with 100nM MitoTracker Red CMXRos (M-7512) (Molecular Probes,

Eugene, OR), a dye taken up by active mitochondria, for 15 minutes at 37oC, and fixed in 3.7%

formaldehyde. Incubation with primary antibody against Mtd was done overnight and secondary

detection was performed as described above for embedded sections.

2.11.6 siRNA Treatment

JEG-3 cells were seeded into a six-well plate at a density of 1x105 cells per well until optimal

confluency for siRNA transfections experiments (>70%) was reached. Cells were then washed

with 1xPBS and 1.5mL of fresh media was added to each well. Subsequently, cells were

transfected with siRNA duplexes (30M) against Mtd and a scramble sequence, purchased from

Ambion (Applied Biosystems, Austin, TX, USA), employing a liposome-based reagent:

lipofactamine2000

(Invitrogen, Carlsband, CA, USA). Sequences were re-suspended in nuclease

free water for a final stock concentration of 50mM. siRNA duplexes and lipofactamine2000

were

first incubated separately for 5 minutes at room temperature, in 0.250mL of OPTI-MEM

(Invitrogen, Carlsband, CA & GIBCO BRL, Grand Island, NY, USA) and then incubated

together for 30 minutes. Treatment was performed over a 48hr time course at standard conditions

and fresh media was added to the cells 6-8hrs following transfection.

63

2.11.7 BrdU Incorporation

BrdU (5-Bromo-2‟-deoxy-uridine) labeling was performed in accordance with the

manufacturer‟s protocol: Roche Applied Sciences (made in: Mannheim, Germany; distributed in

Indianapolis, IN USA). Cells were plated on coverslips and subjected to BrdU (1:1000) for either

20 or 45 minutes following either doxycycline (0.25ug/ml) or Mtd-L siRNA (30uM) treatments.

Cells were then fixed for 20 minutes at -20oC in ethanol fixative (70 ml ethanol, 30 ml of 50mM

glycine, pH to 2.0). Fixed cells were then washed with PBS, blocked with 5% horse serum and

subjected to anti-BrdU antibody for 30 minutes at 37oC. Cells were then washed and 2

o anti-

mouse antibody was applied for 1 hour at room temperature. DAPI staining and mounting were

performed as described for IF staining.

2.11.8 TGF treatment

For experimental procedure cell lines were cultured in either standard conditions (5% CO2 in

95% air) or an atmosphere of 3% O2/92% N2/5% CO2, in the presence or absence of 10 ng/ml of

TGF1 or TGF3 for 24hours.

2.12 Construction of Stable Cell Line Expressing GFP-hMtdL

Human embryonic kidney (HEK) cell line stably expressing GFP-hMtd-L was generated as

follows: plasmid encoding GFP-hMtd-L was stably introduced into Flp-In T-Rex-293 cell line

(Invitrogen as described in the manufacturer‟s protocol. Briefly, the human Mtd-L gene was

amplified from full-length cDNA hMtd-L (Open Biosystems) by PCR using the forward primer

5‟-ggcgcgccagaggtgctgcggcgctcctcg-3‟ and the reverse primer 5‟-cagagagatgacccggatcccg-3‟.

The PCR was digested by AscI/BamHI and cloned into pcDNA5/FRT/TO/GFP (a kind gift from

Dr. Gingras, SLRI at Mount Sinai Hospital, Toronto). The resulting plasmid was verified by

digest and sequencing and finally co-transfected along with pOG44 (Invitrogen) into Flp-In T-

Rex-293 cells to induce a site-specific integration event. Hygromycin-resistant clones were

obtained under 100-ug/ml hygromycin selection. Mtd-L was induced in the stable transfected

cell line with 0.05–2.5 ug/ml of doxycycline.

2.13 Statistical analysis

Statistical analyses were performed using GraphPad Prism 4 software (San Diego, CA). For

comparison of data between multiple groups we used one-way ANOVA Kruskal-Wallis test with

64

Dunns post-hoc test. For comparison between 2 groups we used Mann-Whitney U test.

Significance was defined as P< 0.05. Results are expressed as the mean standard error of the

mean (SE) or box and whisker plots showing medians and interquartile ranges.

65

3 Pro-apoptotic Mtd/Bok Regulates Trophoblast Cell

Proliferation during Human Placental Development and in

Preeclampsia

Note: The content presented in this chapter was first published in Cell Death and Differentiation

(please refer to reference Ray,J.E., Garcia,J., Jurisicova,A., and Caniggia,I. 2009). Contents of

the publication have been reproduced in this data chapter with official permission from the

publisher.

3.1 Abstract

We have previously reported that Mtd/Bok, a pro-apoptotic member of the Bcl-2 family,

regulates human trophoblast apoptosis and that its levels are elevated in severe preeclamptic

pregnancy. Herein we demonstrate that Mtd is also involved in the regulation of proliferation in

normal placentae. Mtd was found in proliferating trophoblast cells during early placental

development, and co-localized with cyclin E1, a G1/S phase cell cycle regulator. The main

isoform of Mtd associated with trophoblast proliferation was Mtd-L, the full length isoform,

which preferentially localized to the nuclear compartment in proliferating cells while during

apoptosis it switched localization to the cytoplasm where it associated with mitochondria.

Antisense specific knock-down of Mtd-L in early first trimester villous explants, as well as loss

and gain of function studies in HEK293 cell line, revealed a direct effect of Mtd-L on cyclin E1

expression and cell cycle progression. We conclude that Mtd-L functions to regulate trophoblast

cell proliferation during early placentation. Of clinical relevance, we hypothesize that the

elevated levels of Mtd found in preeclampsia, may contribute to the increased trophoblast

proliferation accompanying this disorder.

66

3.2 Introduction

In humans, early placental development is defined by an intricate balance between cellular

proliferation, differentiation and death of the trophoblast lineage, the cells forming the placenta

(Smith et al., 1997;Lea et al., 1999;Levy and Nelson, 2000). These cellular events are closely

linked and likely regulated by many of the same molecules.

Early development takes place in a relatively hypoxic environment (~20 mmHg), whereby low

oxygen acts as a key regulator of early trophoblast differentiation (Rodesch et al., 1992;Jaffe et

al., 1997;Burton et al., 1999). In this environment trophoblast proliferation is abundant, whereas

the rate of trophoblast cell death is low (Smith et al., 2000). By 10-12 weeks of gestation, the

oxygen levels increase to ~55 mmHg. This is accompanied by changes in the expression pattern

of a number of gene products (Genbacev et al., 1997;Soleymanlou et al., 2005a), and is

associated with a decrease in trophoblast proliferation and an increased susceptibility to cell

death (Genbacev et al., 1997;Smith et al., 2000). This fine tuning of trophoblast turnover is

governed by a variety of molecules expressed by the placenta, including those comprising the

Bcl-2 family (Ray et al., 2008;Heazell et al., 2008a).

The Bcl-2 family of molecules, classically known for their involvement in the regulation of

apoptosis, include both cell death suppressors (Bcl-2, Bcl-xL, Mcl-1, A1) and cell death inducers

containing either three Bcl-2 homology (BH) domains (Bax, Bak, and Mtd/Bok) or inducers with

only a single (BH3) domain (Hrk, Bim, Bad, Bik, Noxa and Puma). Members of this gene family

act through a complex network of homo- and hetero-dimers with limited specificity. The pro-

apoptotic multi-domain members are believed to regulate apoptosis by forming channels in the

outer membrane of mitochondria leading to the release of pro-apoptogenic factors (Green and

Reed, 1998;Finucane et al., 1999;Soleymanlou et al., 2005b). In recent years it has become

apparent that a number of Bcl-2 family members also regulate cell cycle progression (Bonnefoy-

Berard et al., 2004;Zinkel et al., 2006;Maddika et al., 2007;Ray et al., 2008). Interestingly, while

anti-apoptotic multi domain members slow down progression through the cell cycle, pro-

apoptotic molecules appear to promote cell cycle progression (Brady et al., 1996;Fujise et al.,

2000;Knudson et al., 2001;Bonnefoy-Berard et al., 2004;Jamil et al., 2005;Zinkel et al.,

2006;Maddika et al., 2007).

67

Mtd is a pro-apoptotic Bcl-2 family member that is highly expressed in reproductive tissues (Hsu

et al., 1997;Soleymanlou et al., 2005b). Mtd, is alternatively spliced and encodes three protein

isoforms Mtd-L, Mtd-S, and Mtd-P, with the L and P isoforms predominating in the human

placenta (Hsu and Hsueh, 2000;Soleymanlou et al., 2005b). Similar to Bax, all isoforms of Mtd

contain three BH domains and a transmembrane domain which facilitate pro-apoptotic activity

via mitochondrial depolarization (Soleymanlou et al., 2005b). Previously, we reported that Mtd-

L and Mtd-P expression is high during early placental development (Soleymanlou et al., 2005b).

As this period is characterized by intense trophoblast cell proliferation and little trophoblast cell

death, we hypothesize that Mtd, in addition to its classical role in apoptosis, may have a function

in regulating trophoblast cell proliferation. Interestingly, in vitro studies have shown that the Mtd

promoter can be activated at the G1/S boundary by the E2F1/3 transcription factor (Rodriguez et

al., 2006), providing indirect evidence that Mtd may contribute to regulation of cell cycle

progression. However, the mode by which Mtd exerts its function in the cell cycle remains to be

established.

Herein we report that Mtd is expressed in proliferative trophoblast cells during early placental

development, where it plays a direct role in regulating cyclin E1 expression and promoting G1 to

S phase transition. Furthermore, the dual role of Mtd in apoptosis and proliferation was

associated with a change in sub-cellular localization. Our results indicate that, in addition to its

role in apoptosis, Mtd may be involved in the regulation of trophoblast cell proliferation during

the early stages of human placentation.

3.3 Results

3.3.1 Mtd expression in proliferating trophoblast cells

To determine the pattern of Mtd protein expression in proliferating cells during early gestation,

we performed dual labeled immunofluorescence (IF) analysis with antibodies against Mtd and

Ki67, a common marker of proliferation (Endl et al., 2000) (Figure 3.1). Mtd localization

displayed a unique spatial and temporal pattern of expression. At 5-7 weeks of gestation Mtd was

observed primarily in the cytotrophoblast layer, displaying a prevalent nuclear localization

(Figure 3.1a), whereas weak staining for Mtd was apparent in the syncytiotrophoblast layer

(Figure 3.1a). By 10-13weeks Mtd expression became restricted primarily to the apical border

of the syncytiotrophoblast, with weak nuclear and cytoplasmic expression in the trophoblast

68

Figure 3-1: Mtd expression in proliferating trophoblast cells.

Spatial localization of Mtd and Ki67 in placental sections from the first trimester a: representative early first

trimester (5-8 weeks) floating villous, b: representative late first trimester (9-12 weeks) floating villous, c: early first

trimester anchoring villi (6 weeks), and d: late first trimester anchoring villi (11 weeks). Immunopositivity for Mtd

(green), Ki67 (red), and nuclei labeled with DAPI (blue). Merged images show co-localization of Mtd, Ki67 and

DAPI (overlap of red and blue: pink; overlap of green and blue: light blue; overlap of red, green and blue: white).

Lower panels show boxed region at high magnification. (CT: cytotrophoblast; ST: syncytiotrophoblast; PC:

proximal column; DC: distal column; arrow heads: nuclear positivity, arrows: cytoplasmic positivity). Panel e:

negative controls; first trimester floating and anchoring placental sections (6 weeks) immunostained with

mouse/rabbit IgG.

69

layers (Figure 3.1b). Ki67 expression was restricted to the nuclei of the cytotrophoblast cells and

decreased with advancing gestation (Figure 3.1). Localization of Mtd to Ki67 positive cells

within the cytotrophoblast layer was abundant at 5-7weeks (Figure 3.1a), with 59% of

cytotrophoblast cells expressing both proteins (Table 3.1). However, Mtd expression was not

exclusive to Ki67 positive cells. At 10-13 weeks co-localization of Mtd and Ki67 was less

frequent (Figure 3.1b) with the majority of the cells (57%) being Mtd and Ki67 negative (Table

3.1).

We next investigated whether Mtd localized to Ki67 positive extravillous trophoblast cells

(EVT) forming the anchoring columns. Mtd and Ki67 expression was similarly distributed

throughout the villous column during early first trimester (Figure 3.1c), with co-localization

occurring in 60% of cells (Table 3.1). In contrast, in late first trimester (Figure 3.1d), only 27%

of EVT showed co-localization (Table 3.1). Interestingly, Mtd displayed both nuclear and

cytoplasmic expression at 5-7 weeks (Figure 3.1c), whereas, its localization became

predominantly cytoplasmic by 10-13 weeks (Figure 3.1d).

3.3.2 Mtd localizes to villous trophoblast cells in the G1-phase of the cell cycle

Although Ki67 can be used to indicate cell proliferation it does not discriminate between the

various stages within the cell cycle. To investigate whether Mtd was expressed in the G1 phase of

the cell cycle we tested the co-localization of Mtd with cyclin E1, a cyclin specific to the late G1

phase (Lew et al., 1991;Sherr et al., 2004). Similar to Ki67, cyclin E1 expression was restricted

to the nuclei of cytotrophoblast and EVT within floating villi (Figure 3.2a) and anchoring

columns (data not shown) throughout gestation. Dual labeling of Mtd with cyclin E1 revealed

that Mtd was present during the G1 phase of the cell cycle in early first trimester cytotrophoblast

cells of floating and anchoring villi (Figure 3.2a, data not shown), with co-expression

decreasing in the late first trimester (data not shown). To confirm that Mtd expression was

localized to cytotrophoblast cells during early first trimester, we performed co-expression studies

of Mtd with the cytotrophoblast marker, E-cadherin (Brown et al., 2005)(Figure 3.2b). Of note,

Mtd expression was not restricted to cyclin E1 positive cells; as cells that were presumably

undergoing mitosis, as shown by chromosomal patterning (Figure 3.2c) and Ki67 (Figure 3.2c,

bottom panel), also displayed Mtd expression in the nucleus.

70

Table 3-1 Expression of Ki67 and Mtd in trophoblast cells

Expression of Ki67 and Mtd in trophoblast cells from floating and anchoring villi during first

trimester of gestation (values are reported as percentage)

71

Figure 3-2: Association of Mtd with cyclin E1.

a-b: Spatial localization of Mtd (green) co-localized with a: cyclin E1 (red), or b: E-cadherin (red) in an early first

trimester (6week) floating villous. Nuclei are detected by DAPI (blue). Lower panels show boxed region at high

magnification. a: Arrowheads: co-localization of Mtd with cyclin E1. Upper-right panel; negative control. b:

Arrowheads: representative cells positive for both Mtd and E-cadherin. c: Mtd in mitotic cytotrophoblast cells; Mtd

(green), Ki67 (red) and Top panels: Merged images show co-localization of Mtd, and DAPI (overlap of green and

blue: light blue). Lower panels: Merged images show co-localization of Mtd, Ki67 and DAPI. Middle and right

hand panels show boxed region at high magnification. (CT: cytotrophoblast; S: stroma; ST: syncytiotrophoblast,

arrows: mitotic cells).

72

3.3.3 Mtd expression can occur independently of cell death during early

placentation

We next assessed the association of Mtd expression with the incidence of cell death during the

first trimester using terminal deoxynucleotidyl transferase-mediated dUTP nick-end labeling

(TUNEL), cleaved caspase-3 staining, and analyses of nuclear morphology (Figure 3.3a-d). Cell

death, assessed by these parameters was sporadic and increased with advancing gestation. The

occasional cells that were positive for either TUNEL, cleaved caspase-3, or displayed

fragmented nuclei in early gestation, were found in the stroma (Figure 3.3b), the

syncytiotrophoblast (Figure 3.3a), the syncytial sprouts (Figure 3.3c, upper panel) and the

distal portion of the anchoring villi (Figure 3.3c, lower panel). As expected, Mtd could be

detected in the occasional apoptotic cells as identified by apoptotic blebbing (Figure 3.3c upper

panels), and in the distal portion of the anchoring column (Figure 3.3c, lower panels). No sign

of apoptosis was evident in the cytotrophoblast cells (Figure 3.3).

Interestingly, punctuate expression of Mtd was observed in the nuclei and to a lower extent in the

cytoplasm of proliferating cells (Figure 3.3d) whereas in apoptotic cells Mtd appeared

aggregated and accumulated in the cytoplasm (Figure 3.3c right panels).

3.3.4 Mtd-L is the predominant isoform expressed in proliferative trophoblast

cells

Transcript expression of Mtd-L and Mtd-P, the primary isoforms of Mtd expressed by the

placenta (Soleymanlou et al., 2005b), was examined in proliferating versus non-proliferative

trophoblast cells using Laser Capture Microdissection (LCM), as isoform specific antibodies

toward Mtd-L and Mtd-P were unavailable.

Placental sections from 5-7 and 10-13 weeks were stained for Ki67 to identify proliferative

trophoblast cell populations (data not shown). Adjacent sections were then used to isolate

proliferative EVT, non-proliferative EVT and villous trophoblast cells by LCM (Figure 3.4a).

Both Mtd-L and Mtd-P transcripts were expressed in all three trophoblast populations examined

however; Mtd-L was expressed with Ct (threshold cycle) values ranging between 24-29

(moderate abundance), whereas Mtd-P only exhibited Ct values ranging between 31-35 (very

low abundance) (Figure 3.4b). Expression of Mtd-L in the villous trophoblast layers exhibited

an increase from early to late first trimester and in the anchoring columns Mtd-L mRNA

73

Figure 3-3: Apoptosis in early first trimester placental sections.

a: TUNEL staining in a 6 week placental section. TUNEL (red) DAPI (blue). Arrow: TUNEL positive ST cells. b:

Immunohistochemical staining for cleaved caspase-3 in a 6 week placental section. Cleaved caspase-3 (red) DAPI

(blue). Right panel: higher magnification of the boxed area. Arrow: fragmented pieces of nuclei in the cleaved

caspase-3 positive trophoblast cell. Arrowhead: nuclei of adjacent trophoblast cell showing no sign of apoptosis. c:

Mtd expression in apoptotic trophoblast in syncytial knots (upper panel) and distal EVT (lower panel) from early

first trimester (7 week placenta). Mtd (green), DAPI (blue). Middle and right panels: high magnification of the

boxed areas. Arrow: fragmented pieces of nuclei in the Mtd positive cell. Right panel showing clumped Mtd

expression in the cytoplasm of an apoptotic cells. d: Mtd localization to cytotrophoblast cells that express the

proliferative marker ki67. Mtd (green), Ki67 (red), DAPI (blue). Middle and right panels: high magnification of the

boxed area. Arrowhead: punctuate expression of Mtd. (CT: cytotrophoblast; S: stroma; ST: syncytiotrophoblast;

distal EVT: distal portion of extravillous trophoblast column).

74

Figure 3-4: Mtd isoform mRNA expression in trophoblast subpopulations.

a: Laser capture microdissection of the non-proliferative, distal portion of an anchoring column (7 week placenta).

(PC: proximal column, DC: distal column). b: Expression of Mtd L and P isoforms by qPCR. Threshold cycle was

taken as the point where the curve contacted the dashed line c: Graphical representation of Mtd-L mRNA in villous

(open bars), proliferative EVT in proximal column (hatched bars) and non-proliferative EVT distal column (black

bars) in first trimester placental sections. N=6 for samples from 5-8 weeks, n=5 for samples from 9-12 weeks. Data

did not reach statistical significance P<0.05 Kruskal-Wallis test.

75

expression shifted from the proliferative EVT at 5-7 weeks to the non-proliferative EVT cell

subpopulation at 10-13 weeks; although this did not reach significance (Figure 3.4c).

3.3.5 Mtd isoforms are differentially localized within proliferative JEG-3 cells

Sub-cellular localization of specific Mtd isoforms was investigated in human choriocarcinoma

JEG-3 cell line under normal cell culture conditions. Expression of Mtd-L and Mtd-P isoforms in

this cell line was confirmed by qPCR where Mtd-L was the predominant isoform expressed

(Figure 3.5a). Sub-cellular fractionation demonstrated that Mtd-L was predominantly localized

to the nuclear and light membrane organelle fractions whereas Mtd-P was found in the light

membrane organelle and cytoplasmic fractions (Figure 3.5b).

To determine if Mtd localizes to the mitochondria, we subjected JEG cells to Mitotracker, a

specific mitochondrial tracer dye, and assessed for Mtd co-localization by immunofluorescence

(Figure 3.5c-e). It was confirmed that a subset of Mtd localized to the mitochondria in healthy

(Figure 3.5d), mitotic (Figure 3.5e) and in apoptotic cells (data not shown). To verify that a

similar localization of Mtd is found in vivo we examined first trimester human placental sections

dual labeled with antibodies against Mtd and markers of various cellular organelles. No co-

localization was seen between Mtd and the nuclear envelope, the endoplasmic reticulum, the

golgi apparatus, the cell membrane or lysosomes (data not shown). However, consistent with our

cell line studies, a subset of Mtd co-localized with mitochondria in the cytotrophoblast cells (data

not shown).

3.3.6 SNP-induced apoptosis promotes mitochondrial localization of Mtd in JEG-

3 cells

To determine if Mtd preferentially localized to the mitochondria during apoptosis, we assessed

Mtd localization in JEG-3 cells following apoptotic induction with SNP (Sodium nitroprusside),

a nitric oxide donor recognized to induce apoptosis in the JEG-3 cell line (Soleymanlou et al.,

2007). Cell death and viability were assessed over 48 hours by trypan blue exclusion and MTT

assays (Figure 3.6). In the untreated control group cell proliferation was maximal between 24

and 48 hours (Figure 3.6) with a death rate below 5% [3.86±2.12, 2.94±1.26]. Two days

exposure to 2.5mM and 5mM SNP inhibited proliferation (Figure 3.6) and triggered cell death

[2.5mM: 23.65±4.07% and 5mM: 73.82±6.57]. Nuclear morphology was

76

Figure 3-5: Subcellular localization of Mtd isoforms in JEG-3 cells.

a: mRNA expression of Mtd-L (filled bar) and Mtd-P (hatched bar) in JEG-3 cells grown in standard conditions for

48 hours. b: Fractionation of JEG-3 cells grown in standard conditions for 48 hours. Upper panel: Western blot for

Mtd. Lower panel: verification of cellular fractions: tubulin (cytoplasmic), cytochrome c (mitochondrial), and lamin

A (nuclear membrane). c-e: JEG-3 cell grown in standard conditions for 24 hours labeled with mitotracker (red) and

immuno-stained for Mtd (green), DAPI (blue). Mitotic cells denoted by an asterix. Co-localization of Mtd and

mitochondria: yellow (overlap of red and green) d: JEG-3 cell at high magnification displaying normal nuclear

morphology e: high magnification of JEG-3 cells in mitosis. *P<0.01, Mann Whitney U test.

77

Figure 3-6: Cell viability of SNP treated JEG-3 cells.

a: Percentage of cell death measured by trypan blue exclusion in JEG-3 cells subjected to increasing doses of SNP

(1mM to 10mM) over 24 hours. b: Percentage of cell death measured by trypan blue exclusion in JEG-3 cells

untreated (open bar) and subjected to 2.5mM (hatched bar) and 5mM (filled bar) SNP treatment over 48 hours. c:

Assessment of cell viability by MTT assay in JEG-3 cells subjected to SNP treatment over 48 hours: untreated

(circle), 2.5mM SNP (square unfilled), 5mM SNP (triangle), 10mM SNP (square filled). In all cases bars represent

standard error over three independent experiments.

78

assessed 24 hours following treatment. Mitotic structures were readily seen in untreated

conditions (Figure 3.5c,e, Figure 3.7a), whereas nuclear blebbing was frequently observed in

both SNP treated conditions (Figure 3.7b,c). Mtd was assessed by fluorescence

immunocytochemistry, in JEG-3 cells treated with mitotracker. Mtd could be seen in both the

nuclear and cytoplasmic compartment of cells in the early stages of apoptosis, displaying nuclear

condensation and reformation (Figure 3.7a-c, top right panels). With progressive degree of

apoptosis, evidenced by increased nuclear transformation and blebbing, Mtd expression became

less nuclear and predominantly cytoplasmic where it appeared aggregated and localized to

mitochondria (Figure 3.7b, lower panel. 3.7c, middle panel). At late stages Mtd could be

detected in the cytoplasm but its localization to mitochondria could not be found as the

mitochondria were no longer capable of uptaking the tracer dye (Figure 3.7c, bottom panel).

3.3.7 Inhibition of Mtd-L suppresses cyclin E1 expression

In order to determine the functional significance of Mtd in trophoblast cell cycle during the first

trimester we evaluated the consequences of inhibiting Mtd-L on cyclin E1 expression in first

trimester human placental explants using an antisense knockdown approach. This approach has

been previously used in our laboratory with a knockdown efficiency in explants of 40-60%

(Soleymanlou et al., 2005b). The inhibition of Mtd-L resulted in a 31.5% and 31.3% reduction of

both cyclin E1 mRNA and protein expression, respectively (Figure 3.8a,b). In addition this was

accompanied by a marked decrease in PCNA expression, a marker of S phase (Figure 3.8b,

middle panel). Interestingly, knockdown of Mtd-L had no effect on cyclin E1 expression in

explants from late first trimester of gestation (data not shown).

The Caniggia lab has previously reported that overexpression of Mtd-L and Mtd-P results in

cellular apoptosis (Soleymanlou et al., 2005b). In order to establish a role for Mtd in cellular

proliferation, we used a doxycycline-inducible Mtd-L expression system in HEK293 cells,

generated in the lab; which allowed for controlled Mtd-L expression (Figure 3.8c). Induction of

Mtd-L using 0.25ug/ml doxycycline over 36hours did not affect the rate of cell death as

determined by trypan blue exclusion (data not shown) and it was associated with 32% increase

in cyclin E1 protein expression (Figure 3.8c). This was accompanied by an increase in BrdU

incorporation by these cells (Figure 3.8d). Conversely, an siRNA-mediated reduction of Mtd-L

79

Figure 3-7: Localization of Mtd to mitochondria in apoptotic JEG-3 cells

JEG-3 cells were left untreated (a), treated with 2.5mM SNP (b) or 5mM SNP (c) for 24 hours. Cells were labeled

with mitotracker (red) and immuno-stained for Mtd (green). Co-localization of Mtd and mitochondria: yellow

(overlap of red and green). Arrowhead denotes apoptotic cells; Star denotes area of apoptotic cells; Asterix denote

mitotic cells. 100x images show apoptotic cells representative from each treatment group. a: occasional apoptotic

cell in untreated group displaying early stages of apoptosis and co-localization of Mtd with mitotracker. b: cells

showing the occasional mitotic structure and increased levels of cell death c: numerous apoptotic cells displaying a

range from early (condensed, blebbing nuclei, top right panel) to blebbing and fragmented nuclei (middle panel) and

apoptotic bodies (lower panel).

80

Figure 3-8: The effect of Mtd-L interference on cyclin E1.

a,b: Expression of cyclin E1 (mRNA: n=4; protein: n=4) in early first trimester explants treated with Mtd-L Sense

(S, non-silencing) oligonucleotides or Mtd-L antisense oligonucleotides (AS, silencing) or untreated (C) in 3% O2

for 72 hours. a: Relative transcript levels of cyclin E1 as detected by qPCR and corrected against 18S. *P<0.05,

Mann Whitney U test. b: Expression levels of cyclin E1 protein (upper panel) and PCNA protein (middle panel).

Confirmation of equal loading by Ponceau detection of total proteins (lower bands). c-d: Induction of GFP-hMtdL in

Flp-In T-Rex-293 cell line with increasing concentrations of doxycycline (0.05, 0.25, 0.5, 1.25, and 2.5ug/ml) over a

36hour period (n=3). c: Top panel: Level of hMtdL expression detected by GFP. Middle panel: Expression of cyclin

E1 and bottom panel, control actin. HEY ovarian cancer cell lysate was included as a negative control for GFP and a

positive control for cyclin E1. d: BrdU incorporation in hMtdL Flp-In T-Rex-293 cell line following treatment with

or without 0.25ug/ml doxycycline over a 36hour period. e-g: knockdown of MtdL in HEK293 cells using siRNA

strategy (n=3). e: Fold changes in MtdL transcript levels relative to scrambled sequence (SS) as detected by qPCR.

Statistical significance was assessed by Kruskal-Wallis or Mann Whitney U test; a,b,c: P<0.05. f: Expression levels

of cyclin E1 protein and actin. g: BrdU incorporation in HEK293 following knockdown of MtdL using 30uM

siRNA for 48hours. Negative control; Immunostaining with mouse IgG in HEK cells subjected to BrdU.

81

(75%) in HEK293 cells lead to a decrease in both cyclin E1 protein expression (41%) and BrdU

incorporation (Figure 3.8f, g).

3.4 Discussion

In recent years it has become apparent that members of the Bcl-2 family are involved in

regulating cell fate in a variety of systems (Bonnefoy-Berard et al., 2004;Zinkel et al.,

2006;Maddika et al., 2007). Data presented herein demonstrate that Mtd, a pro-apoptotic

molecule of the Bcl-2 family, is involved in the regulation of both proliferation and cell death in

the human placenta in physiological and pathological conditions. In particular we demonstrate

that 1) Mtd is expressed in proliferating trophoblast cells during early placental development,

where it acts on the G1 phase of the cell cycle by directly regulating the expression of cyclin E1,

2) that Mtd-L is the isoform responsible for this effect on cell proliferation, and 3) that Mtd

localizes to the nuclear compartment in proliferating cells while during apoptosis it switches

localization to the cytoplasm where it interacts with mitochondria.

Both anti-apoptotic and pro-apoptotic members of the Bcl-2 family have been shown to

participate directly in cell cycle regulation independent of their apoptotic function (Bonnefoy-

Berard et al., 2004;Zinkel et al., 2006;Maddika et al., 2007). Anti-apoptotic members, Bcl-2,

BCL-xL, BCL-w, and Mcl-1, have been shown to have inhibitory effects on passage through the

cell cycle whereas the pro-apoptotic member, Bax, promotes cell proliferation by conferring cell

cycle advancement (Brady et al., 1996;Knudson et al., 2001;Zinkel et al., 2006). Although the

Mtd promoter has been found to be activated at the G1/S boundary in vitro (Rodriguez et al.,

2006), we present for the first time, direct evidence that the Mtd protein is expressed in cycling

cells in vivo. In the human placenta the expression of Bax associates predominantly with the

apoptotic index (De Falco et al., 2001), suggesting that Bax may not be involved in trophoblast

cell cycle regulation. This underscores a unique function for Mtd as a pro-apoptotic regulator of

cell cycle progression in the human placenta.

Our study found that nuclear expression of Mtd closely associated with the proliferative index

throughout human placental development. In the early first trimester Mtd localized primarily to

the nucleus in Ki67- and cyclin E1- positive cytotrophoblast cells that displayed normal nuclear

morphology with no sign of apoptosis, whereas past the 9th

week of gestation, as the percentage

of proliferative cells decreased, expression of Mtd switched to the apical border of the

82

syncytiotrophoblast layer. The effect of Mtd on cyclin E1 was also associated with this temporal

regulation, occurring only in early first trimester explants. Nonetheless, the apoptotic role of Mtd

has been shown to be maintained past the first trimester (Soleymanlou et al., 2005b). Taken

together, these data suggest that a change in Mtd function may take place late in the first

trimester, away from a cell cycle regulatory role, and that this is accompanied by a change in

sub-cellular localization. This switch in localization and function is likely regulated by the

increase in oxygen concentration experienced by the placenta at this time. This is supported by

previous studies that have found hypoxia response elements in the promoter region of Mtd (Gao

et al., 2005;Soleymanlou et al., 2005b), and have shown Mtd to be decreased in higher oxygen

conditions (Gao et al., 2005;Soleymanlou et al., 2005b).

Since the inhibition of Mtd leads to a decrease in cyclin E1 we hypothesize that Mtd may

function during the G1/S transition. In accordance with our observations a recent study

performed with a mouse fibroblast cell line showed expression of the Mtd transcript to be

increased at mid to late G1 phase, following overexpression of G1/S phase transition factors

(Rodriguez et al., 2006). Moreover the decrease in PCNA and BrdU incorporation following Mtd

knockdown and increased BrdU incorporation following doxycycline-induced Mtd

overexpression, indicates that the effect on cyclin E1 expression results in altered cell cycle

progression to the S phase. In addition, we also observed Mtd expression in mitotic cells and in

Ki67 negative cells. It is possible that like Mcl-1, which functions at both the G0/G1 and at the

G2/M border (Fujise et al., 2000;Jamil et al., 2005), Mtd may play different roles at distinct

phases of the cell cycle(Fujise et al., 2000;Jamil et al., 2005).

Our data also suggest that the role of Mtd in cell fate is likely dependent on conformation and

location of the protein. This mode of regulation has previously been shown for Bax, a Bcl-2

family member structurally similar to Mtd (Hsu et al., 1997;Inohara et al., 1998). Whereas Bax is

cytoplasmic and monomeric in proliferating cells it‟s apoptotic function depends upon its

oligomerization and translocation to the mitochondria (Wolter et al., 1997;Antonsson et al.,

2000;Antonsson et al., 2001;Dejean et al., 2005). Similarly, we observed a diffuse, punctuate

expression pattern of Mtd in proliferative trophoblast cells, whereas cells undergoing apoptosis

displayed a clumped or aggregated pattern of Mtd expression, consistent with oligomerization.

Furthermore, Mtd has previously been reported to form oligomers under apoptotic stimuli in

HEK293 cells (Gao et al., 2005).

83

In addition, Mtd expression was both cytoplasmic and nuclear in proliferating cells, whereas its

expression became cytoplasmic and localized to the mitochondria in cells undergoing apoptosis.

Mtd has been shown to interact with the exportin Crm1 (Bartholomeusz et al., 2006), supporting

the idea that Mtd can travel between the nuclear and the cytoplasmic compartments to exert its

cellular function. It is therefore plausible that in trophoblast cells, low levels of Mtd remain

monomeric and locate to the nucleus, where their function is linked to cell cycle regulation,

whereas cytoplasmic accumulation of Mtd promotes oligomerization, localization to the

mitochondria, and a functional switch towards its apoptotic role.

Interestingly, Mtd was also observed in a sub-set of mitochondria in proliferating JEG-3 cells.

This may be explained by previous work showing that in MCF-7 cells, Mtd loosely associates

with the mitochondria and that upon apoptotic stimulation, Mtd becomes tightly integrated into

the mitochondrial membrane (Gao et al., 2005).

Mtd is expressed as three isoforms in total human placental lysate, the principle isoforms being

Mtd-L and Mtd-P (Soleymanlou et al., 2005b). We postulate that the Mtd-L isoform has a dual

role in both cell proliferation and death while Mtd-P may have primarily a “killing” role. This is

supported by our fractionation studies that revealed Mtd-L to be the predominant isoform

expressed in the nuclear compartment of proliferating JEG-3 cells, where it would have direct

access to interaction with cell cycle regulating molecules, and by our knockdown and

overexpression experiments where Mtd-L was seen to have a direct effect on cyclin E1

expression. In contrast, both Mtd-L and Mtd-P were located in the mitochondrial fraction of

JEG-3 cells and their overexpression has been shown to result in apoptosis (Soleymanlou et al.,

2005b).

Based on a putative working model (Figure 3.9) we postulate that during early placental

development, Mtd-L is upregulated by the low oxygen environment, and that it localizes to the

nucleus of the cytotrophoblast cells where it promotes proliferation by aiding in the G1 to S

transition. Introduction of apoptotic stimuli may lead to cytoplasmic accumulation of Mtd, and

result in Mtd oligomerization and interaction with the mitochondria, thus producing a pro-

apoptotic response. In conclusion, Mtd appears to play an important role in the proliferative and

apoptotic pathways that mediate trophoblast cell fate. Of clinical relevance, improper regulation

of Mtd, as seen in preeclampsia, may play a role in altering the homeostasis of trophoblast

84

layers, contributing to the increased rate of proliferation and apoptosis associated with this

pathology.

85

Figure 3-9: Putative model of Mtd function in the trophoblast

Under low oxygen conditions, as seen in early placental development, Mtd remains monomeric and localizes to the

nucleus of cytotrophoblast cells where it promotes proliferation by aiding in the G1 to S transition. Under conditions

of oxidative stress, Mtd accumulates in the cytoplasm where it interacts with the mitochondria. Mtd pore formation

in the mitochondria leads to release of apoptogenic molecules in to the cytoplasm, thus activating of the apoptotic

cascade and cell death.

86

4 Altered trophoblast proliferation in preeclampsia is

associated with increased cyclin E1expression and

abnormal regulation of the cell cycle inhibitor p27

4.1 ABSTRACT

While knowledge concerning molecular pathways regulating human placental development has

increased dramatically, regulation of trophoblast cell cycle remains poorly understood. Recently

we reported that Mtd, a pro-apoptotic member of the Bcl-2 family, promotes the expression of

cyclin E1, an activator of G1/S phase transition of the cell cycle. Moreover, we have previously

shown Mtd to be elevated in the preeclamptic pathology where it is involved in apoptosis of the

trophoblast. Herein we demonstrate that Mtd is also expressed in proliferating trophoblast cells

in preeclampsia. In addition we investigate the regulation of cyclin E1, and the CDK inhibitor,

p27. During normal development levels of cyclin E1 decreased with gestational age as p27

increased. In contrast, the protein levels of both cyclin E1 and p27 were elevated in severe early

onset preeclampsia (SPE), compared to age matched controls (AMC). In addition, the majority of

p27 was phosphorylated at the Ser10 site and localized predominantly to the cytoplasm of cyclin

E1 positive cytotrophoblast cells in SPE. This phenomenon was specific to early onset severe

preeclampsia, as placentae from late onset preeclamptic (LPE) and intra uterine growth restricted

(IUGR) pregnancies did not show differences in expression of either cyclin E1 or p27. We have

previously reported on both increased TGF3 and a hypoxic environment in preeclamptic

placentae, factors that drive trophoblast differentiation. In this study it was found that the

combined treatment of TGF3 and 3%O2 lead to an increase of cyclin E1 and p27 expression in

cultured placental explants and the JEG choriocarcinoma cell line. Our data suggest that the

elevated levels of Mtd and TGF3, in addition to the hypoxic environment in preeclamptic

placentae, may contribute to the increased trophoblast proliferation accompanying early onset

severe preeclampsia.

87

4.2 Introduction

Preeclampsia, a serious pregnancy related disorder currently faced by the obstetrician, affects 3-

8% of pregnancies, and if left untreated can lead to fetal and maternal death and morbidity

(Roberts and Cooper, 2001). Currently, diagnosis of this disorder is based on the onset of the

maternal symptoms, including the sudden onset of maternal hypertension and proteinurea, which

typically do not manifest until the second or third trimester once the disease has already been

established. In addition, the etiology of preeclampsia remains unknown. It is well recognized that

the placenta plays a central role in its pathogenesis, as its removal at delivery is the only

treatment known to resolve the maternal symptoms. It is therefore important that effort be put

forward to decipher the underlying cellular and molecular defects of this placental disorder so

that early diagnosis and treatment can be improved.

Physiologically, the placenta has a likeness to a controlled cancer; in the early stages of

pregnancy its growth is exponential and the tissue is highly invasive. Extravillous trophoblast

cells differentiate, from proliferative to invasive cells, that invade the uterine wall to obtain

access to oxygen and nutrients for the developing fetus (Graham et al., 1991;Kurman, 1991b).

Meanwhile the nutrient exchanging syncytiotrophoblast, bathed in the maternal blood, is

constantly sloughed off and replenished through fusion of the underlying proliferative

cytotrophoblast cells (Kaufmann, 1982). The extensive proliferation of cytotrophoblast cells in

the early stages of pregnancy allows for considerable expansion of the organ, facilitating the

increase in surface area and the nutrient absorbing capacity of the syncytiotrophoblast.

In cases of preeclampsia however, the placenta exhibits excessive trophoblast cell turnover, a

process involving increased trophoblast proliferation, fusion and extrusion (DiFederico et al.,

1999;Allaire et al., 2000;Leung et al., 2001;Huppertz et al., 2004). Consequently, this leads to

release of excessive placental debris in to the maternal circulation (Johansen et al., 1999;Redman

and Sargent, 2000), causing maternal endothelial dysfunction and the maternal symptoms of

preeclampsia. In addition, preeclampsia has been associated with a deficiency of the extra villous

trophoblast cells to differentiate into cells capable of adequately infiltrating the uterine arteries.

This results in poor placental perfusion, and insufficient nutrient delivery to the fetus. Of note,

preeclampsia can manifest in either early (<34 wks) or late pregnancy (>34wks) and in many

cases is accompanied by intrauterine growth restriction of the fetus. However, the molecular

88

differences defining the individual types of preeclampsia are still being uncovered and are

currently not fully understood.

Studies have reported that trophoblast cells in preeclamptic placentae are arrested to an immature

hyperproliferative phenotype, that may account for the increased cell turnover and poorly

invasive properties (Arnholdt et al., 1991;Brown et al., 2005). The precise mechanisms leading

to the increased proliferation and turnover in preeclampsia however, have yet to be identified.

The Caniggia lab has reported that Mtd-L and Mtd-P expression levels are significantly

increased in preeclamptic placentae and that this is associated with increased trophoblast cell

death (Soleymanlou et al., 2005b). Moreover, we have previously identified a role for Mtd-L in

the regulation of cyclin E1, an activator of cyclin-dependent kinase-2 (CDK2) and promoter of

the G1 to S transition, in normal placental development (Ray et al., 2009). However, the

relationship between Mtd expression and the hyperproliferative nature of preeclamptic

trophoblast cells remains unexplored. During normal placentation the regulation of trophoblast

proliferation and differentiation is also mediated in part by the oxygen status of the

microenvironment as well as the associated changes in growth factor expression including

members of the TGF family (Jaffe et al., 1997;Genbacev et al., 1997;Caniggia et al.,

1999;Caniggia et al., 2000;Caniggia et al., 2002). Importantly, preeclampsia is associated with

placental hypoxia and a hyperactive TGF pathway (Gerretsen et al., 1981;Caniggia et al.,

1999;Hung et al., 2002;Soleymanlou et al., 2005a). Additionally, the Caniggia lab has previously

shown that cyclin E1 is upregulated by TGF3 (unpublished data).

Herein we investigate whether Mtd is associated with trophoblast proliferation in placentae from

severe early onset preeclampsia. In addition we examine the regulation of cyclin E1 and p27, a

classic CDK2 inhibitor, with respect to the proliferative capacity of the trophoblast cells in the

human placenta. Here we reveal that Mtd was expressed in proliferative trophoblast cells in

placentae from pregnancies complicated with severe early onset preeclampsia. In addition we

show that this disorder was associated with increased expression of both cyclin E1 and p27,

which may lead to excessive trophoblast cell production and possibly improper trophoblast

differentiation. Furthermore, we demonstrate that the low oxygenation and increased TGF

associated with preeclampsia may potentiate altered cell cycling through increased cyclin E1

expression and cytoplasmic localization of p27.

89

4.3 Results

4.3.1 Mtd expression in proliferating trophoblast cells in preeclampsia

We have previously reported that Mtd expression is elevated in placentae from severe

preeclamptic pregnancies and that this is associated with increased trophoblast cell death,

characteristic of this disease (Soleymanlou et al., 2005b). To determine whether the increase in

Mtd was also associated with the hyper-proliferative nature of the trophoblast cells we tested

whether Mtd was expressed in Ki67 and cyclin E1 positive cells in placental samples from

severe early onset preeclampsia. In addition we tested whether Ki67 and cyclin E1 levels were

elevated in placentae from severe early onset preeclampsia (PE) relative to age matched (AMC)

and term controls (TC). As anticipated, preeclampsia displayed greater levels of Ki67 (% of Ki67

positive trophoblast cells, PE: 8.6% vs AMC: 3.7% and TC: 1.7%; p< 0.05), cyclin E1 (% of

cyclin E1 positive trophoblast cells, PE: 13% vs AMC: 6.3% and TC: 2.5%; p< 0.05), and Mtd

compared to age-matched and term controls (Figure 4.1a,b, Figure 4.2a,b). Furthermore, dual

labeling of Mtd with Ki67 and cyclin E1 was evident to a greater extent in preeclamptic samples

compared to either age-matched (Figure 4.1a,b, Figure 4.2a,b) or term controls (data not

shown). Interestingly, nuclear expression of Mtd in preeclamptic samples was predominant in

Ki67 or cyclin E1 positive cells whereas in the syncytial knots the expression was mainly

cytoplasmic. In contrast, comparable levels of Mtd were seen in both the nuclear and

cytoplasmic regions of age-matched (Figure 4.1a,b and Figure 4.2a,b bottom panels) and term

control sections (data not shown). Co-expression of Mtd with E-cadherin confirmed that

cytotrophoblast cells were the predominant sites of nuclear Mtd expression in preeclamptic

placentae (Figure 4.2c).

4.3.2 Cyclin E1 and the CDK inhibitor p27 show opposing expression during

normal placentation

Since Mtd appears to impact the cell cycle at the level of cyclin E1, we performed experiments to

gain insight into how cyclin E1 may be regulated and contribute to normal placental

development. In general, the cell cycle is initiated by expression of G1 phase cyclins in response

to a variety of pro-proliferative cues in the environment (Sherr et al., 2004). Since the placenta

experiences a physiological change in oxygenation (Jauniaux et al., 2000) and growth factor

milieu in the late first trimester of normal pregnancy (Caniggia et al., 2000), we evaluated the

90

Figure 4-1: Localization of Mtd to proliferative cells in preeclamptic placentae.

a: Spatial localization of Mtd-L and Ki67 in placental tissue from preeclamptic (PE) and age-matched control

(AMC) placentae. Mtd (green), Ki67 (red) and DAPI (blue). Top panels: floating villi from a preeclamptic placenta

at 29 weeks of gestation, bottom panels: floating villi from an age-matched control placenta (34wks). Left hand

panels: 20X magnification; middle and right panels: 100X magnification. Merged panels show co-localization

(yellow/white) of Mtd with Ki67 with nuclei detected by DAPI staining. CT: cytotrophoblast cells; SK: syncytial

knots; arrows: co-localization of Mtd and Ki67 in the nuclear compartment of cells in the trophoblast layer. b:

quantification of the percentage of trophoblast cells positive for Ki67 in PE (n=5), AMC (n=5) and term controls

(TC) (n=5). Statistical significance was assessed by Kruskal-Wallis with Dunn‟s post hoc test; a,b,c: P<0.05.

91

Figure 4-2: Localization of Mtd to cyclin E1 positive cytotrophoblast cells in preeclamptic placentae.

a: Spatial localization of Mtd and Cyclin E1 in floating villi from a PE placenta at 26 weeks of gestation (top pane)

and from an AMC placenta of 32wks (bottom panel). Mtd (green), Cyclin E1 (red) and DAPI (blue). Arrowhead: co-

localization of Mtd and cyclin E1 in the nuclear compartment of cells in the trophoblast layer. b: quantification of

percentage of trophoblast cells positive for Cyclin E1 in PE (n=5), AMC (n=5) and term controls (TC) (n=5).

Statistical significance was assessed by Kruskal-Wallis with Dunn‟s post hoc test; a,b,c: P<0.05. c: Spatial

localization of Mtd and the cytotrophoblast marker E-cadherin in placental tissue from preeclamptic (32 weeks) and

age-matched control placentae (32 weeks). Mtd (green), E-cadherin (red) and DAPI (blue). Left hand panels: 20X

magnification; middle and right panels: 100X magnification. Open arrowhead: cytotrophoblast cell co-expressing

Mtd and E-cadherin. (CT: cytotrophoblast cells; SK: syncytial knots);

92

protein expression of cyclin E1 in total placental lysates from gestational windows of 5-8weeks,

9-14 weeks and term (Figure 4.3). Cyclin E1 protein expression was found to be significantly

increased during early first trimester compared to term; paralleling the mitotic index previously

published for the placenta (Figure 4.3a). Increased expression of cyclin E1 was associated with

an elevated expression of its binding partner, cyclin dependent kinase 2 (CDK2) (Figure 4.3c),

as well as PCNA, an S phase marker (Figure 4.3d). CDK inhibitors are important negative

regulators of the cell cycle. In particular, the cip/kip inhibitor p27, was found to preferentially

bind to cyclin E/CDK2, inhibiting the CDK2 activity by interfering with the catalytic cleft

(Russo et al., 1996). Hence, we next assessed the expression of p27 in the placenta over

gestation. In contrast to cyclin E1, protein levels of p27 increased significantly with each

advancing window of gestational age (Figure 4.3b).

We next assessed the cellular localization of cyclin E1 and p27 during normal placental

development by dual labeled florescence immunohistochemistry. Trophoblast cells in the

floating and anchoring villi expressed both cyclin E1 and p27 (Figure 4.4). In the trophoblast

layer of the floating villi, expression of cyclin E1 was restricted to the nuclei of the

cytotrophoblast cells throughout gestation (Figure 4.4a,b). Additionally cytoplasmic expression

of cyclin E1 could be observed in the endothelial cells of the villous vessels (Figure 4.4a,b).

Expression of p27 occurred in both the cytotrophoblast and syncytiotrophoblast layer as well as

in cells of the stroma in the floating villi throughout development (Figure 4.4a,b). However,

during the early first trimester expression of p27 was predominantly nuclear (Figure 4.4a) while

in the late first trimester and term, its expression became both cytoplasmic and nuclear (Figure

4.4b, data not shown).

In the anchoring columns both cyclin E1 and p27 were expressed in the extra villous trophoblast

(EVT) throughout the column (Figure 4.4c,d). Although the EVT of the proximal site of the

anchoring column exhibit proliferative capabilities, the intensity of cyclin E1 expression in the

late first trimester was increased toward the distal portion of the column (Figure 4.4d), where

the cells have been shown to be more differentiated. In contrast to cyclin E1, expression of p27

expression remained consistent throughout the column in both early and late first trimester

samples (Figure 4.4c,d). Co-localization of nuclear p27 with cyclin E1 in the EVT was reduced

in the late first trimester compared to the early first trimester as evidenced by the decrease in

93

Figure 4-3: Cyclin E1 and p27 expression during placental development

a-d: Representative Western blot and densitometric analysis for cyclin E1(a), p27(b) CDK2 (c), and PCNA (d) over

gestation; (cyclin E1: 5-8 weeks, n=14; 9-14 weeks, n=19; term, n=11) (p27: 5-8 weeks, n=20; 9-14 weeks, n=26;

term, n=17) (CDK2: 5-8 weeks, n=15; 9-14 weeks, n=14; term, n=10) (PCNA: 5-8 weeks, n=9; 9-14 weeks, n=9;

term, n=7). 5-8weeks (filled bars), 9-14 weeks (unfilled bars) and term (hatched bars); Data were normalized against

actin. Statistical significance was determined as P<0.05 Kruskal-Wallis followed by Dunn‟s test.

94

Figure 4-4: Spatial localization of cyclin E1 and p27 in human placentae from first trimester and term

a,b: Immunohistochemical dual staining of p27-green and cyclin E1-red in an early first trimester (5weeks), and

term (37week) floating placental villi. Nuclei are visualized by DAPI labeled chromatin (blue). Middle and right

panels: high magnification of the boxed areas. Arrow: cytotrophoblast cell expressing both cyclin E1 and p27

Arrowhead: endothelial cells expressing nuclear p27 and cytoplasmic cyclin E1 (CT: cytotrophoblast; ST:

syncytiotrophoblast; S: stroma; V: vessel). c,d: spatial localization of p27-green, and cyclin E1-red in extra villous

trophoblast of (c) an early first trimester (5week) and (d) a late first trimester (11 week) anchoring villi. (S: stroma;

PC: proximal column; DC: distal column). e: Immunoprecipitation of p27 in total placental tissue lysates over

gestation followed by immunoblotting for cyclin E1 and 27. Input lane (lane 10) is 30ug of total term placental

lysate (same sample as lane 9).

95

combined immuno-reactivity (yellow) (Figure 4.4c,d bottom panels). Immunoprecipitation

studies confirmed the interaction between cyclin E1 and p27 throughout placental development

and further demonstrated that the interaction between p27 and cyclin E1 decreased from early

first trimester to late first trimester and term (Figure 4.4e).

4.3.3 Expression of cyclin E1 and p27 is altered in severe early onset

preeclamptic placentae compared to age matched and term controls

Since we observed an increase in cyclin E1 expression in preeclampsia and since trophoblast

cells of preeclamptic placentae are poorly differentiated and maintained in a proliferative

immature state, we examined the overall expression pattern of cyclin E1 and p27 in this

pathology. In line with our immunohistochemical analysis, preeclamptic placentae displayed

significantly up-regulated levels of cyclin E1 protein compared to age matched and term controls

(P<0.05, P<0.01 respectively) with a 2.5 fold increase in protein level compared to term (Figure

4.5b). Similarly, fluorescence immunohistochemical analysis revealed that placentae from

preeclamptic patients display elevated levels of cyclin E1 expression in the cytotrophoblast cells

of the floating villi compared to age matched controls (Figure 4.5e). Cyclins have a half life of

20-30 mins with their protein levels typically paralleling that of their transcript. Surprisingly,

cyclin E1 mRNA was not significantly different in preeclamptic samples compared to controls

(Figure 4.5a), suggesting that regulation of cyclin E1 may be altered by posttranslational

regulation in preeclampsia.

We next assessed whether the increase in trophoblast proliferation seen in preeclampsia was

associated with a decrease in expression of the cell cycle inhibitor p27, similar to that seen in

early placental development. Unexpectedly, levels of p27 were elevated in preeclamptic

placentae compared to age matched and normal term controls (P<0.01, P<0.001) (Figure 4.5d).

In comparison to age matched control sections, p27 expression in PE was predominantly

cytoplasmic in the trophoblast layer and mainly nuclear in the endothelial cells surrounding the

villous vessels (Figure 4.5f). In addition, co-localization studies revealed that p27 expression

was primarily cytoplasmic in cyclin E1 positive PE cytotrophoblast cells (Figure 4.6b).

However, immunoprecipitation of p27 resulted in an increased association of p27 with cyclin E1

in preeclamptic samples compared to controls, a finding that may reflect the overall abundance

of p27 in the preeclamptic samples (Figure 4.6a). No interaction was evident between p27 and

96

Figure 4-5: mRNA and protein expression of cyclin E1 and p27 in placentae from pregnancies associated with

severe early onset preeclampsia.

a,c: qRT-PCR analysis of cyclin E1 and p27 mRNA from total placental tissue lysates from severe early-onset

preeclampsia (PE, filled bars n=10), normotensive age-matched (AMC, unfilled bars, n=7), and term control

samples (hatched bars, n=9). Data were normalized against expression of 18S ribosomal RNA using the well

established 2-CT

formula; fold changes are relative to term control levels. b,d: Representative Western blot and

densitometric analysis for cyclin E1 and p27 protein in placental lysates from severe early onset preeclampsia, age-

matched and term control samples (cyclin E1: PE n=14, AMC n=10, TC n=10) (p27: PE n=28, AMC n=17, TC

n=19). Samples are corrected by actin and normalized to term controls. Statistical significance was determined as

P<0.05, Kruskal-Wallis with Dunn‟s post hoc test. e,f: Spatial localization of cyclin E1 (red) or p27 (green) in

severe early onset preeclamptic, and age-matched control placental sections. Nuclei are visualized by DAPI labeled

chromatin (blue). Right hand panels: high magnification of the boxed areas. e: Arrow: expression of cyclin E1 in

cytotrophoblast cells; Arrowhead; syncytiotrophoblast cells negative for cyclin E1 f: Fetal blood cells are visible by

autofluorescence in red. Arrow: expression of p27 in cytotrophoblast and syncytiotrophoblast cells; Arrowhead; p27

expression in endothelial cells (CT: cytotrophoblast; ST: syncytiotrophoblast; SK: syncytial knot; V: vessel).

97

Figure 4-6: Cyclin E1 and p27 interaction in preeclamptic pregnancies

a: Immunoprecipitation of p27 in severe early onset preeclampsia (PE), age matched controls (AMC) and Term total

placental tissue lysates followed by immunoblotting for cyclin E1, cyclin D1, and p27. Input lane (lane 10) is 30ug

of total term placental lysate (from same sample as lane 9). b: Spatial localization of cyclin E1 (red) with total p27

(green) in severe early onset preeclamptic (PE-34weeks), and age-matched control (AMC-32week) placental

sections. Nuclei are visualized by DAPI labeled chromatin (blue). Right hand and lower panels: high magnification

of the boxed area with and without DAPI visualization. Arrow: cytotrophoblast cells expressing both cyclin E1 and

p27 in a preeclamptic placentae; Arrowhead cytotrophoblast cell expressing cyclin E1 and low level of p27 (CT:

cytotrophoblast; SK: syncytial knot).

98

cyclin D1. Interestingly, no change in p27 transcription could be observed in preeclamptic vs age

matched or term control samples (Figure 4.5c), suggesting that the increase of p27 seen in

preeclampsia may occur at the level of post translational regulation. No differences in cyclin E1

or p27 expression were detected for mRNA or protein between age matched and term controls

(Figure 4.5/4.6, and data not shown).

4.3.4 Post translation regulation of p27 is altered in preeclampsia

Localization and functional regulation of p27 is highly regulated through phosphorylation.

Phosphorylation of p27 at the Ser10 residue can lead to its export from the nuclear compartment

into the cytoplasm, where p27 can no longer inhibit cyclin E1-CDK2 mediated G1/S transition

(Besson et al., 2006). In addition, phosphorylation of p27 at the Thr157 or Thr198 sites result in

the protein‟s stabilization and persistence in the cytoplasm, which may account for the increased

levels of total p27 seen in preeclampsia (Fujita et al., 2002;Liang et al., 2007). Using a p27 Ser10

phospho-specific antibody (p-p27 Ser10) we determined that phosphorylation of p27 at the Ser10

site was upregulated in placentae of preeclamptic patients compared to age matched or term

controls (Figure 4.7a). In addition, immunofluorescent studies confirmed that Ser10

phosphorylation of p27 did confer cytoplasmic retention of the protein, and it revealed that the

majority of the cytotrophoblast cells expressing p-p27 Ser10 were cyclin E1 positive (Figure

4.7d). Interestingly, our studies demonstrated no appreciable difference in phosphorylation at

either the Thr157 or Thr198 p27 residue in preeclampsia compared to control samples (Figure

4.7b,c).

4.3.5 Regulation of cyclin E1 and p27 are altered in several placental

pathologies

To examine whether the observed increase in cyclin E1 and p27 was specific to the severe early

onset form of preeclampsia we assessed cyclin E1 and p27 protein expression in placentae from

two related placental pathologies; late onset preeclampsia (LPE) and intra uterine growth

restricted (IUGR) pregnancies. In contrast to the early onset severe form of preeclampsia, neither

LPE nor IUGR samples displayed significantly elevated levels of cyclin E1 or p27 expression

compared to their respective age matched controls (Figures 4.8a,b; 4.9a,b). However in LPE

there was a trend toward an increase in p27 compared to term (Figure 4.8b) and cyclin E1 levels

were increased in preterm IUGR cases relative to the term control samples (Figure 4.9a).

99

Figure 4-7: Phosphorylation of p27 in severe early onset preeclampsia

a: Representative Western blot and densitometric analysis of p27 phosphorylated at Ser10, normalized to Ponceau

(upper graph), and relative to total p27 levels (lower graph), in placental lysates from severe early-onset

preeclampsia (PE, filled bars, n=14), normotensive age-matched (AMC, unfilled bars, n=10), and term control

samples (hatched bars, n=10). b,c: Representative Western blot and densitometric analysis of p27 protein

phosphorylated at (b) Thr157 and (c) Thr198 in placental lysates from severe early-onset preeclampsia (PE, filled

bars), normotensive age-matched (AMC, unfilled bars), and term control samples (hatched bars). Protein expression

was normalized to actin for densitometric analysis. d: Spatial localization of cyclin E1 (red) with p-p27 Ser-10

(green), in sever early onset preeclamptic (30weeks), and age-matched control (24week) placental sections. Nuclei

are visualized by DAPI labeled chromatin (blue). Right hand panels: high magnification of the boxed areas.

Arrowhead: Cytotrophoblast cells expressing nuclear cyclin E1 and cytoplasmic phospho-p27-ser-10; Arrow;

trophoblast cells negative for cyclin E1 and p-p27 Ser10. Statistical significance was determined as P<0.05 Kruskal-

Wallis followed by Dunn‟s test.

100

Figure 4-8: Expression of cyclin E1 and p27 in pregnancies complicated by late onset preeclampsia.

a-c: Representative Western blot and densitometric analysis of (a) cyclin E1, (b) total p27, and (c) p-p27 Ser10,

normalized to Ponceau (upper graph), and relative to total p27 levels (lower graph), in placental lysates from late-

onset preeclampsia (LPE, filled bars, n=13) and term control samples (TC, unfilled bars, n=11). Protein expression

was normalized to Ponceau for densitometric analysis. Statistical significance was determined as P<0.05 Mann

Whitney U test. d,e: Spatial localization of cyclin E1 (red) with (d) total p27 (green) or (e) p-p27 Ser10 (green), in

late onset preeclamptic (36week), and term (38,42 week) placental sections. Nuclei are visualized by DAPI labeled

chromatin (blue). Middle and right hand panels: high magnification of the boxed areas with or without DAPI

visualization. Arrowhead: Trophoblast cells expressing nuclear p27; Arrow; trophoblast cells expressing cyclin E1.

(SK: syncytial knot; V: vessel; S: stroma).

101

Figure 4-9: Expression of cyclin E1 and p27 in pregnancies exhibiting IUGR

a-c: Representative Western blot and densitometric analysis of (a) cyclin E (b) total p27, and (c) p-p27 Ser10 in

placental lysates from pregnancies complicated by intra uterine growth restriction (IUGR, filled bars, n=14),

normotensive age-matched (AMC, unfilled bars, n=8), and term control samples (TC, hatched bars, n=10). Protein

expression was normalized to actin expression for densitometric analysis. P<0.05 Kruskal-Wallis with Dunn‟s test.

d,e: Spatial localization of cyclin E1 (red) with (d) total p27 (green) or (e) p-p27 Ser10 (green), in IUGR (30,32

week), and age matched control (31,32 week) placental sections. Nuclei are visualized by DAPI labeled chromatin

(blue). Middle and right hand panels: high magnification of the boxed areas with or without DAPI. Arrowhead:

Trophoblast cells expressing cytoplasmic p27; Arrow; trophoblast cells expressing cyclin E1. (V: vessel; S: stroma).

102

Interestingly, Immuno-localization studies of cyclin E1 exposed a more prominent pattern of

expression in the vasculature of both the LPE and IUGR pathologies compared to their

respective controls (Figure 4.8d,e; 4.9d,e). In addition, p27 expression was localized to a greater

extent in the nuclear cell compartment of the trophoblast cells in LPE compared to term

placentae (Figure 4.8d). In contrast, sections from IUGR placentae displayed a comparable

pattern of p27 to their AMC controls (Figure 4.9d). No observable differences in p-p27 Ser10

were detected by western blot or immunofluorescent analysis between LPE and term, or IUGR

and age matched control samples (Figure 4.8c,e; 4.9c,e).

4.3.6 Phosphorylation of p27 is increased in the early stages of normal placental

development

To understand the relevance of p27 phosphorylation at Ser10 in severe early onset preeclampsia,

we examined the expression of p-p27 Ser10 in normal placental development. Although no

overall abundance of p-p27 Ser10 was observed in placental lysates throughout gestation (Figure

4.10a upper graph), the ratio of p-p27 Ser10 to total p27 was found to be significantly higher

during early first trimester compared to term (Figure 4.10a lower graph).

Dual labeled florescence immunohistochemistry of p-p27 Ser10 with cyclin E1 revealed that

during normal placental development p-p27 Ser10 was predominantly expressed in the

cytoplasm of cyclin E1 positive cytotrophoblast of the floating villi (Figure 4.10b) and the

cyclin E1 positive extra villous trophoblast cells of the distal anchoring column (Figure 4.10c).

Furthermore, in late first trimester samples, expression of p-p27 Ser10 was reduced in the

cytotrophoblast and its expression became evident to a greater extent in the villous stroma

(Figure 4.10b).

4.3.7 TGF influences cyclin E1 and p27 expression in villous explants cultured

under varying oxygen conditions

We have previously shown that trophoblast differentiation in the late first trimester is driven by

the change in oxygen status and associated decrease in TGF3 expression (Everett and

MacDonald, 1979;Rodesch et al., 1992;Caniggia et al., 2000). We therefore tested whether

treatment with TGF1 or TGF3 under various oxygen conditions had an effect on the protein

103

Figure 4-10: Expression Ser10 phosphorylated p27during placental development

a: Representative Western blot and densitometric analysis for p27 phosphorylated at Ser10 normalized to -actin

(upper graph), and relative to total p27 levels (lower graph), in placental lysates over gestation; (5-8 weeks, n=12; 9-

14 weeks, n=12; term, n=10). 5-8weeks (filled bars), 9-14 weeks (hatched bars) and term (unfilled bars); Statistical

significance was determined as P<0.01 Kruskal-Wallis with Dunn‟s test. b,c: Spatial localization of cyclin E1 (red)

with p-p27 Ser10 (green), in an early first trimester (5week), and late first trimester (12week) floating placental villi

(b) and anchoring villi (c). Nuclei are visualized by DAPI labeled chromatin (blue). Middle and right panels: high

magnification of the boxed areas. Arrow: cytotrophoblast cell expressing both cyclin E1 and p-p27 Ser10.

Arrowhead: cytotrophoblast cells expressing cyclin E1 alone (CT: cytotrophoblast; ST: syncytiotrophoblast; S:

stroma; PC: proximal column; DC: distal column).

104

expression of cyclin E1 or p27 in the trophoblast using a first trimester explants model (Figure

4.11).

Interestingly, neither cyclin E1 nor p27 expression was increased when cultured in 3%

oxygenation conditions compared to 20%. Furthermore, neither treatment with TGF1 nor

TGF3 resulted in an increase in the expression of cyclin E1 or p27 under 20% oxygenation. In

contrast, the combined treatment of TGF in 3% oxygen culture conditions resulted in a trend

toward an increase in both cyclin E1 and p27 protein expression compared to untreated controls,

however this trend did not reach significance (Figure 4.11 top and middle panels). Interestingly

explants treated with either TGF1 or TGF3 displayed decreased levels of p-p27 Ser10 at each

oxygen level tested, when compared to untreated controls (Figure 4.11 bottom panels).

4.3.8 TGF influences cyclin E1 and p27 expression in JEG-3 choriocarcinoma

cell line cultured under varying oxygen conditions

JEG-3 choriocarcinoma trophoblast cell lines were employed as a cell model to further test the

effect of oxygen and TGF on the expression of cyclin E1 and p27 in the trophoblast. Similar to

our explant model, cells cultured at 3% oxygen and treated with TGF1 or TGF3 displayed a

trend toward an increase of both p27 and cyclin E1 compared to cells untreated (Figure 4.12a

top and middle panels). Interestingly, similar effects were seen between cells cultured at 3%

and those cultured at 20% oxygen (Figure 4.12). In contrast to the explant model, JEG

choriocarcinoma cells displayed increased phosphorylation of p27 at Ser10 when treated with

either TGF1 or TGF3 (Figure 4.12a bottom panels).

Subcellular localization of cyclin E1 and total p27 was assessed in TGF3 treated JEG cells by

dual labeled fluorescence immunohistochemistry. Compared to untreated control cells, those

treated with TGF3 displayed an observable increase in cyclin E1 positive cells and an increased

incidence of mitotic figures (Figure 4.12b). Furthermore, cyclin E1 was detected in association

with mitotic figures (ii), in structures consistent with the appearance of spindles (iv) and in nuclei

of cells with average and excessively large nuclei (i,iv). P27 showed an increase in both

cytoplasmic and nuclear expression upon TGF3 treatment (Figure 4.12b)

105

Figure 4-11: Effect of TGF on cyclin E1 and p27 expression in first trimester villous explants exposed to

various oxygen conditions.

a: Representative Western blot analysis of cyclin E1, p27, and p27 phosphorylated at Ser10 from early first trimester

placental explants treated with or without TGF1 or TG3, cultured in 3%, 8% or 20% oxygen condition. b:

Protein expression was normalized to -actin for densitometric analysis. N=3 for each treatment group. P<0.05

Kruskal-Wallis with by Dunn‟s test. Data did not reach statistical significance.

106

Figure 4-12: Effect of TGF on cyclin E1 and p27 expression and localization in JEG-3 choriocarcinoma

trophoblast cell line exposed to different oxygen conditions

a: Representative Western blots and densitometric analysis of cyclin E1, p27, and p27 phosphorylated at Ser10 in

JEG-3 trophoblast choriocarcinoma cells treated with or without TGF1 or TGF3, cultured in 3%, or 20% oxygen

condition. Three independent experiments were run in triplicate. Protein expression was normalized to -actin for

densitometric analysis (n=3) P<0.05 Kruskal-Wallis followed by Dunn‟s test. Data did not reach statistical

significance. b: subcellular localization of cyclin E1 (red) and p27 (green) in JEG-3 trophoblast choriocarcinoma

cells treated with or without TGF3, cultured in 3%, or 20% oxygen conditions. Nuclei are visualized by DAPI

labeled chromatin (blue). Middle and right panels: high magnification of the boxed areas.

107

4.4 Discussion

Altered rates of trophoblast proliferation and the subsequent increase in cell turnover are factors

underling severe early onset preeclampsia. Surprisingly, little is known about the underlying

mechanisms leading to the hyperproliferative phenotype of the trophoblast in this pathology.

Recently we reported that Mtd is increased in preeclampsia and that Mtd has a positive effect on

cell cycle progression. Herein we established that Mtd is expressed in proliferative cells in the

cytotrophoblast layer and that levels of cyclin E1 and p27 protein are also elevated in placentae

from severe early onset preeclampsia. Secondly, that the inhibitory function of p27 is hampered

due to increased phosphorylation at its Ser10 site resulting in its nuclear export. We also

determined that the increase in TGF and oxidative stress associated with the preeclamptic

disorder may alter the levels of cyclin E1 and p27, and lastly that severe early onset preeclampsia

displays a molecular profile distinct from late onset preeclampsia or IUGR.

We have previously shown that both the Mtd-L and Mtd-P isoforms are increased in

preeclamptic placental tissue compared to age-matched and term controls (Soleymanlou et al.,

2005b). Preeclampsia is a placental disorder, characterized by hyper-proliferation of the

trophoblast cells and accompanied by excessive apoptosis and syncytial shedding. We therefore

hypothesized that Mtd may in part contribute to the increase of both cellular apoptosis and cell

proliferation in this pathology. We have previously shown that in preeclampsia, Mtd is

associated with mitochondrial depolarization and an increase in apoptosis (Soleymanlou et al.,

2005b), and in the current study we demonstrate co-localization of Mtd to the Ki67- and cyclin

E1-positive cytotrophoblast cells in preeclampsia, suggesting that the abundance of Mtd may

also contribute to the hyper-proliferative phenotype. This is further supported by our data

showing that in preeclampsia, the hyperproliferative phenotype of the trophoblast cells is

associated with increased Ki67 and cyclin E1 expression levels. It is likely that Mtd works in

concert with a variety of cell cycle regulating molecules including the Cip/Kip family of

inhibitors and the Notch proteins, which have also been found to be overexpressed in

preeclampsia (reviewed in Crocker and Heazell 2008)(Heazell et al., 2008a).

During normal placentation, the percentage of proliferative trophoblast cells has been shown to

decrease towards term, as the placenta prepares for the final stages of pregnancy (Olvera et al.,

2001). Our data demonstrated that this was associated with a decrease in expression of cell cycle

108

advancing molecules cyclin E1, CDK2, and PCNA as well as an increase in the expression of the

cell cycle inhibitor, p27. Consistent with previous reports, we found that cyclin E1 was highly

expressed in the mitotically active cytotrophoblast layer and in the EVT of the anchoring

columns (DeLoia et al., 1997;Bamberger et al., 1999;Olvera et al., 2001). As cyclin E1 acts as

an important checkpoint required for the G1 to S transition during the cell cycle, it is likely that

the expression of cyclin E1 in the cytotrophoblast cells functions to promote cell proliferation

required for the extensive growth of this organ in the first trimester. Interaction between p27 and

cyclin E/CDK complexes has been shown to inhibit cyclinE/CDK2 activation thereby inhibiting

cell cycle progression (Sherr et al., 1995). The observed localization of p27 to the cyclin E1

positive trophoblast cells as well as the interaction of p27 with cyclin E1 at early gestation

therefore suggest that p27 plays a key role in regulating the activity of cyclinE1-CDK2 during

early placental development. Studies have also shown that p27 is elevated in quiescent cells and

that it is involved in the differentiation of a number of cell types (Sgambato et al., 2000;Besson

et al., 2008). It is therefore possible that p27 may function to arrest the cell cycle in the

cytotrophoblast as well as maintain quiescence in the syncytium. This idea is consistent with our

observation that p27 localized to nuclei of the proliferative cytotrophoblast layer as well as

nuclei of the adjacent non-proliferative syncytium. In support of our findings, McKenzie et al,

found that in vitro, differentiation from CT to ST coincided with a decrease in cyclin E1 and an

increase in p27 expression (McKenzie et al., 1998).

In our study, expression of cyclin E1 and p27 was also observed in the endothelium of the villous

vasculature, suggesting that these two molecules affect cell types at multiple levels of

placentation. Similarly, cyclin E knockout mice have been shown to display defects in yolk sac

vascularization, suggesting that cyclin E plays an important role in vasculogenesis (Parisi et al.,

2003). The cytoplasmic localization of cyclin E1 in the endothelial cells suggests that the role of

cyclin E1 may be independent of its role on G1-S phase progression in human placental

vasculature. Alternatively, the cytoplasmic localization may reflect a functional role in

centrosome duplication (Jackman et al., 2002). In addition to villous expression, cyclin E1 was

highly expressed in the more differentiated, non-proliferating cells of the distal columns. This

suggests that in the EVT, cyclin E1 may function in a role alternative to promoting cell

proliferation. This concept has been exemplified in the cyclin E1/E2 double knock out murine

model, which was found to be embryonic lethal due to defects in placental giant cells, the murine

109

equivalent of human EVT (Geng et al., 2003;Parisi et al., 2003). Alike human EVT, giant cells

are migratory and are believed to function in maternal spiral artery remodeling during

implantation, and in hormone and cytokine production (Cross et al., 1994). In giant cells cyclin E

was shown to play an important role in endoreduplication, a process which allows for multiple

rounds of DNA synthesis to occur in the absence of cell division (Geng et al., 2003;Parisi et al.,

2003). Geng et al postulate that during endoreduplication cyclin E may aid in priming the

chromosomes for replication prior to S phase entry, a function independent of its classical role in

CDK activation. Similar to the murine model, invasive EVT in the human placenta also display

increased DNA content (referred to by Weier et al as aneuploidy), arising from an unknown but

comparable process (Weier et al., 2005). It is therefore plausible that cyclin E1 may have a

similar function in human trophoblast aneuploidy to that of endoreduplication in the murine

model, a line of study that warrants further investigation. In addition Ser10 phosphorylated p27

was also found in the cytoplasm of EVT cells in the distal portion of the anchoring columns. This

may reflect the role of p27 in cell migration, when localized to the cytoplasm (McAllister et al.,

2003). Hence, the wide range of regulatory roles attributed to cyclin E1 and p27 in the placenta

demonstrate their importance in placental development.

It is widely known that perturbation in the expression or function of various cyclins or CDK

inhibitors is linked to several pathologies associated with hyper-proliferation, including

numerous oncogenic diseases. It is therefore plausible that altered expression of cyclin E1 or p27

may result in pathology of the human placenta. This is supported by the cyclin E1/E2 double

knockout murine model that results in lethality due to placental defects as well as a handful of

studies based on Ki67 staining and BrdU incorporation that suggest that trophoblast cells in

preeclamptic placentae are arrested to an immature hyperproliferative phenotype (Arnholdt et al.,

1991;Redline et al., 1995;Geng et al., 2003;Parisi et al., 2003). Moreover, we have demonstrated

previously and in this study, that preeclampsia is associated with an increased incidence of Ki67

and cyclin E1 expression and that the cyclin E1 levels in preeclampsia were similar to that seen

in early development, supporting the theory that PE trophoblast cells are arrested to an immature

state (Ray et al., 2009).

The hyperproliferative phenotype of the preeclamptic trophoblast is an underlying factor

contributing to preeclampsia. The vascular endothelial dysfunction and a systemic inflammatory

response associated with preeclampsia are believed to result from the excessive debris (placental

110

origin) and/or an enhanced sensitivity of the maternal vascular endothelium to a normal

circulating load (maternal origin), the former being associated with severe early onset

preeclampsia and the latter being associated with late onset preeclampsia (Johansen et al.,

1999;Redman et al., 2000). This theory has been supported by the presence of increased

placental material, including syncytial microvillous membrane particles, trophoblast specific

protein cytokeratin, fetal proteins and free fetal DNA, in the maternal circulation of preeclamptic

women, even prior to the onset of the classical symptoms. As well, it has been shown that

placental debris can directly contribute to the inflammatory response of endothelial

cells(Johansen et al., 1999;Redman et al., 2000;Zhong et al., 2001;Levine et al., 2004).

Elevated expression of p27 in severe early onset preeclampsia was unexpected, as

hyperproliferative disorders are often associated with decreased levels of p27. Interestingly, a

number of cancers, including breast, ovarian, colon and oesophageal cancer have also been found

to express increased levels of p27, while maintaining a hyperproliferative state (Besson et al.,

2008). In these cases p27 was seen to be translocated into the cytoplasm of the cell, where its

inhibitory role on cell cycle progression was hampered. Similarly, we observed that in placentae

from preeclamptic pregnancies p27 was predominantly cytoplasmic in cyclin E1 positive cells.

Although p27 can be transcriptionally regulated, the main mechanism of regulation occurs post-

translationally through phosphorylation at a number of serine and threonine sites (Vervoorts and

Luscher, 2008). Moreover, whereas no alteration in p27 mRNA was seen in the preeclamptic

samples compared to controls, these samples displayed significantly increased levels of p27

Ser10 phosphorylation. Phosphorylation at Ser10 has been shown to stabilize p27 by directing

p27 out of the nucleus and in to the cytoplasm through interaction with the exportin CRM1

(Rodier et al., 2001;Connor et al., 2003). This translocation effectively prevents its

phosphorylation at the Thr187 residue and the subsequent proteosomal degradation in the

nucleus (Vlach et al., 1997). Therefore, the cytoplasmic localization and increased expression of

p27 in preeclampsia likely result from the increased phosphorylation of p27 at the Ser10 residue.

Further stabilization of p27 in the cytoplasm can occur through its phosphorylation at Thr157 or

Thr198 or its incorporation into a protein complex such as the cyclinD1-CDK4/6 complex

(Polyak et al., 1994;Fujita et al., 2002;Liang et al., 2007). However, no significant increase in

p27 phosphorylation at either the Thr157 or Thr198 sites was found, nor did p27 interact with

cyclin D1 in the preeclamptic samples.

111

Previous data have shown trophoblast proliferation to increase under low oxygen tension in vitro

and in pregnancy complications involving placental hypoxia, such as maternal anemia and

preeclampsia(Genbacev et al., 1997;Caniggia et al., 2000;Huppertz et al., 2003;Myatt, 2006).

However, our studies in placental explants and JEG cells suggested that low oxygen alone was

insufficient to cause an increase in cyclin E1 or p27.

Our data indicate that it is more likely that the upregulation of cyclin E1 and p27 expression seen

in severe early onset preeclampsia result from the combined effect of altered TGF

(transforming growth factor ), a family of growth factors, and oxidative stress associated with

the disorder. We have previously shown that placental hypoxia increases the expression of TGF

and that TGFs regulate cell proliferation, differentiation, migration and invasion in a number of

cell types including those in the placenta (Polyak et al., 1994;Irving and Lala, 1995;Massague,

1998;Caniggia et al., 1999;Ietta et al., 2006). In addition, the significance of the TGF3 pathway

in preeclampsia has been exemplified by the increase in soluble endoglin levels (a co-receptor of

TGF) and increased activation of the Smad pathway (downstream target of TGF) (Venkatesha

et al., 2006). Moreover, this theory is supported by our in vitro explant and cell line studies that

displayed increased cyclin E1 and p27 expression when treated with TGF under low oxygen

conditions.

In contrast to severe early onset preeclampsia, late onset preeclampsia and IUGR, showed little

appreciable difference in the level of cyclin E1 or p27 expression, compared to their respective

controls. However, placentae from late PE displayed a trend toward an increase in p27 which

may function to maintain proper regulation over cell proliferation in this pathology. This was

supported by the predominant localization of p27 to the nuclei of cyclin E1 positive cells in late

onset preeclampsia. Furthermore, unlike severe early onset preeclampsia that is instigated by the

increase in trophoblast turnover, late preeclampsia is believed to be a maternal rather than

placental disorder. There has been conflicting data as to whether IUGR is a hyperproliferative

disorder (Smith et al., 1998;Chen et al., 2002;Gurel et al., 2003;Jeschke et al., 2006). In contrast

to Jeschke et al 2006, who found the proliferation marker Ki67 to be decreased in IUGR, our

data show a trend toward an increase in cyclin E1 expression supporting the theory that the

IUGR placenta is hyperproliferative. Although late preeclampsia and IUGR are closely related to

severe early onset preeclampsia, sharing similar risk factors and perinatal outcomes, our data

112

support the theory that these pathologies are physiologically different disorders (Villar et al.,

2006;Romero et al., 2008).

Proper trophoblast proliferation is vital for a healthy pregnancy. Insufficient trophoblast

proliferation and invasion during placentation can result in poor implantation and pregnancy

loss, or result in syndromes such as preeclampsia, and IUGR. It is therefore critical that we

understand the molecular basis governing cell cycle regulation in the placenta, so that we can

continue to uncover the mechanisms underlying these disorders. A greater understanding of the

trophoblast cell cycle, and the changes that occur in placental pathology, will inevitably aid in

the discovery of future therapeutic and diagnostic strategies.

113

5 Dual role for Mtd in trophoblast proliferation and apoptosis in

molar pathology

5.1 Abstract

Complete hydatidiform molar pregnancy (CHM) is a devastating condition whereby placental

tissue develops in the absence of a fetus. Although it is established that molar pregnancy is

characterized by excessive trophoblast proliferation and death, the molecular basis governing its

pathogenesis is largely unknown and has been based primarily on descriptive analysis of molar

tissue from the first trimester. We have previously shown that Mtd, a pro-apoptotic member of

the Bcl-2 family, functions to promote both cell death and proliferation in the human placenta.

The objective of this study was to examine the expression of Mtd in the context of cell death and

proliferation in molar pregnancy using a second trimester twin model consisting of a molar and a

genetically normal co-twin placenta, unbiased by environmental conditions. Mtd was elevated in

the molar tissue compared to the co-existing twin placentae where it was associated with

increased TUNEL positive apoptotic cells as well as increased Ki67 and cyclin E1 positive

proliferative trophoblast cells. In addition, placentae from molar pregnancy exhibited decreased

expression of cyclin D1 and D3 protein despite the high level of cyclin E1 present. Cyclin E

function is inhibited by both p21 and p27 cell cycle inhibitors. Interestingly only p27 was

decreased in molar samples whereas p21 expression was significantly increased in comparison with

their co-twin and age-matched control twin sets.

We conclude that molar placentae from second trimester twin pregnancies are associated with a

disruption in the expression of G1 phase cell cycle regulating molecules. The excessive

trophoblast proliferation in these cases is likely due to elevated levels of cyclin E1 and is not

reliant on the presence of D type cyclins. In addition, the increased trophoblast proliferation may

be due to a decreased inhibition by p27, but not p21. These alterations are unrelated to

environmental conditions as placentae from their twin counterpart were unaffected.

114

5.2 Introduction

Complete molar pregnancy, a devastating trophoblastic disease of androgenic chromosomal

origin, is typically characterized by a developing placenta in the absence of a fetus (Slim et al.,

2007). Once identified these pregnancies are therefore immediately terminated. This placental

disorder has a high rate of incidence in Asia occurring in 1:200 pregnant women while it

manifests in 1:1500 Caucasians (Seckl et al., 2000;Steigrad, 2003;Tham et al., 2003) However

the incidence of molar pregnancy is on the rise. Moreover, in many cases, complete molar

pregnancy will persist following evacuation, with 3-33 % of the cases developing into

choriocarcinoma requiring chemotherapy (Kurman, 1991a;Mazur et al., 1994). The malignant

transformation of molar pregnancy is thought to arise from the abnormally high proliferative and

sometimes invasive nature of the trophoblast cells, typical of the disease.

First trimester molar placentae have been shown to exhibit exuberant trophoblast cell

proliferation and excessive cell death in both the stromal and trophoblastic regions. Interestingly

the excessive proliferation found in molar tissue has been associated with elevated markers of

proliferation including Ki67 and cyclin E1 (Kale et al., 2001;Olvera et al., 2001), similar to that

seen in preeclampsia (Ray et al., 2009), and increased apoptosis has been assessed by TUNEL

(Qiao et al., 1998;Chiu et al., 2001). These studies however have been mostly descriptive, and

restricted primarily to immunohistochemical analysis. Absence of the cell cycle inhibitor, p57,

due to paternal imprinting has remained one of the most identifiable features of the complete

mole pathology (McConnell et al., 2009a;Kipp et al., 2010). However it has been shown that

molar tissue can develop in the presence of p57, as seen in two unique cases of molar pregnancy

that expressed p57 due to retention of a single maternal chromosome (Fisher et al.,

2004;McConnell et al., 2009b). It is therefore important to understand the cell cycle regulating

molecules, other than p57, that contribute to the disease. Thus far however, the molecular

alterations contributing to the pathogenesis of this disorder are not fully understood. As such, the

study of molecules involved in trophoblast cell fate in molar pregnancy will provide further

understanding of how improper placentation develops in this pathology, and may provide insight

towards future treatment strategies.

Although the Bcl-2 family of molecules have been shown to play significant roles in trophoblast

cell fate (Ray et al., 2008), few studies have reported on their expression in molar pregnancies

115

(Qiao et al., 1998;Fong et al., 2005;Fong et al., 2006). Studies have found the ratio between pro-

apoptoic Bax and anti-apoptotic Bcl-2 to be elevated in compete molar tissue while others have

suggested a role for anti-apoptotic Mcl-1 in molar pathogenesis (Qiao et al., 1998;Fong et al.,

2005).

Mtd is a pro-apoptotic member of the Bcl-2 family, expressed primarily in tissues of

reproductive origin (Hsu et al., 1997;Soleymanlou et al., 2005b). In preeclampsia, a common

placental disorder complicating 3-5% of all pregnancies (Roberts et al., 2001), we found Mtd to

be elevated in the placenta and have a role in both the apoptotic and proliferative trophoblastic

features of the disease (Soleymanlou et al., 2005b;Ray et al., 2009). In apoptotic cells Mtd was

associated with mitochondrial depolarization and induction of caspase-3 activity, whereas in

proliferative cells Mtd was seen to effect cyclin E1 expression (Soleymanlou et al., 2005b;Ray et

al., 2009). Interestingly, molar pregnancies are associated with a high incidence of preeclampsia

(Soto-Wright et al., 1995;Koga et al., 2009) and may therefore exhibit overlapping dysregulatory

mechanisms with preeclampsia, including increased Mtd expression. However, no study to date

has addressed the expression or function of Mtd in molar pregnancy, despite its high abundance

in the placenta. We therefore postulate that Mtd may be elevated in molar tissue and contribute

to the abnormal balance of apoptosis and proliferation associated with molar pathogenesis.

In normal clinical circumstances the identification of complete molar tissue occurs early on and

results in immediate termination of the pregnancy. Hence, studies examining the molecules

involved in regulating trophoblast cell fate in complete molar pregnancy have been restricted to

tissue of the first trimester. In this chapter, we report on two rare cases of molar pregnancy,

consisting of a complete molar placenta with a co-existing twin; the molar tissue having

developed adjacent to the independent genetically normal placenta of a live fetus. These cases

are unique, as unlike typical moles that are terminated early, the pregnancies were carried on till

delivery which occurred preterm. Therefore, they allow us to perform molecular studies in molar

placental tissue that has been permitted to develop into the second trimester. In addition, these

cases provide twin placental controls, grown in an identical uterine environment for which to

compare the molecular findings.

Herein, we examine the expression of Mtd in the context of cell death and proliferation in two

twin molar cases. As we have shown Mtd to be involved in cell proliferation and apoptosis in

116

normal and preeclamptic placentation, we studied the expression of Mtd in conjunction with

markers of apoptosis, as well as molecules involved in regulating the G1 phase of the cell cycle.

Importantly, comparisons were made between the molar tissue samples and the placentae of their

co-existing twins, as well as between twin controls from normal preterm pregnancies.

5.3 Results

5.3.1 Two cases of a twin pregnancy with a complete hydatidiform mole and

coexistent twin fetus

A twin gestation consisting of a complete hydatidiform mole and a co-existing twin fetus is a

rare event and presents extreme clinical risks. Often these pregnancies develop preeclampsia and

require preterm delivery (Soto-Wright et al., 1995;Koga et al., 2009;Massardier et al., 2009). In

addition they are also at increased risk of additional pregnancy-related complications including

the development of persistent gestational trophoblastic disease (GTD) (Massardier et al., 2009).

Case 1

Patient one, age 37, carried a twin pregnancy with complete hydatidiform mole. It was her first

pregnancy. During her pregnancy she exhibited hyperthyroidism and mild pregnancy induced

hypertension, but she did not develop preeclampsia. She delivered by vaginal delivery at 23

weeks due to chorioamnionitis and sepsis. A male fetus of 485g was born preterm and was not

viable (Table 2.2).

Case 2 (Figure 5.1)

Patient two, age 33, also carried a twin pregnancy with complete hydatidiform mole. It was her

fifth pregnancy and first incidence of molar pregnancy. She presented with severe preeclampsia

at 24 weeks and developed hyperthyroidism. Her pregnancy resulted in neonatal death of a 350g

female fetus at 25 weeks. The fetus was diagnosed as being preterm and growth restricted.

Patient two chose to have a full hysterectomy immediately following the pregnancy (Table 2.2).

In both cases of molar pregnancy the molar placentae demonstrated classic morphological,

anatomical, and histopathological signs of complete mole pathology. These included enlarged

villi with prominent swelling (Figure 5.1b,c), extensive circumferential trophoblastic

117

Figure 5-1 Morphologic characteristics of placentae from the mole and its co-existing twin

a: Macroscopic appearance of the placentae from case 2, with molar placenta left of the hatched line and genetically

normal placenta with gestational membrane and umbilical cord associated with twin fetus at right. b: magnification

of the molar placenta with the swollen cystic villi bulging from beneath the membrane cover (arrows) c-f: Histologic

sections of the mole and twin placenta from case 2 stained with H&E c,d: section of placenta from the mole and the

co-twin presented at equal magnification, Molar villi (left panel) and multiple small mature villi of normal placenta

(right panel). e,f: high magnification of molar villi. Asterisk: dilated cystic villi. HT: hyperproliferative trophoblastic

layer. Arrow, thinned trophoblastic layer.

118

hyperplasia (Figure 5.1e), as well as acellular cistern formation in the central portion of the villi

(Figure 5.1c,e,f) (Shih and Kurman, 2002). The molar villi also displayed no sign of mature

placental vasculature, a common characteristic of molar pathology (Slim et al., 2007;Kim et al.,

2009). In addition, both cases exhibited areas of the enlarged villi with a thinned layer of

trophoblast cells, a feature noted in molar pathology (Figure 5.1f). Both pregnancies were

severely compromised; ending preterm, and resulting in fetal death. The molecular finding from

the molar tissue was therefore compared to placentae from normal preterm (age-matched) twin

pregnancies in addition to the co-existing twins of the moles, to control for any abnormality

resulting in the placenta of the co-existing twin fetus.

5.3.2 Second trimester complete molar placentae display increased trophoblast

proliferation and apoptosis

Since increased trophoblast proliferation and apoptosis have been reported in complete

hydatidiform molar tissue from singleton pregnancies terminated in the first trimester, we first

set out to determine if second trimester molar tissue associated with a co-existent twin fetus

would show a similar phenotype. Proliferation was assessed by immunohistochemistry using an

antibody towards Ki67, a commonly used marker of cell proliferation (Endl et al., 2000) (Figure

5.2a-f).

In both molar twin cases (case 1 and case 2) the molar placenta exhibited an extreme level of

proliferation compared to both its co-existing twin and the control twin sets (Figure 5.2a-f). The

excessive Ki67 positive staining in the molar placenta was found in the trophoblast cells in both

the continuous trophoblast layers and the focal areas of cellular hyperplasia (Figure 5.2a,c). In

tissue sections from the genetically normal placenta from case 2, and from normal control twin

sets, Ki67 was expressed predominantly in the cytotrophoblast cells and in the occasional

mesenchymal cell (Figure 5.2d-f). Unexpectedly, the genetically normal placenta in the first

molar twin set was devoid of Ki67 detection (Figure 5.2b). No obvious differences in Ki67

expression were observed between twin placentae within the same pregnancy in the control twin

sets or between sets of control twins.

To determine if the two second trimester molar cases displayed excessive apoptosis,

characteristic of first trimester moles, we assessed the level of terminal deoxynucleotidyl

transferase-mediated dUTP nick-end labeling (TUNEL) reactivity (Figure 5.3a-f). In contrast to

119

Figure 5-2 Proliferative assessment of placentae from the mole and its co-existing twin

a-f: Immunolocalization of Ki67 in molar tissue (top, middle left), the placenta of its co-existing twin (top, middle

right), and in control twins (bottom panels). Brown staining represents positive Ki67 immunoreactivity.

120

Figure 5-3 Apoptotic assessment of placentae from the mole and its co-existing twin by TUNEL staining

a-f: Labeling of apoptotic cells by enzymatic detection of DNA fragmentation using terminal transferase TUNEL

method in molar tissue (top, middle left), and in the placenta of its co-existing twin (top, middle right) and in control

twins (bottom panels). Brown staining represents positive TUNEL reactivity.

121

the control twin sets, an increased number of apoptotic cells were found in both cases of molar

placentae, occurring mainly in the trophoblastic cells associated with the area of hyperplasia and

in the stromal region of the chorionic villi (Figure 5.3a,c). Limited apoptosis was detectable in

the control twins tested or in the genetically normal placenta form case 2, where the occasional

TUNEL positive cell was observed in the trophoblastic layer or syncytial knots (Figure 5.3d-f).

Interestingly, in the first molar case, an abundance of apoptosis was observed by TUNEL

reactivity in the genetically normal placentae of the co-twin (Figure 5.3b).

5.3.3 Pro-apoptotic Mtd is elevated in the molar placenta compared to its co-

existing twin and it is associated with apoptotic cells in the trophoblastic

and stromal areas

To determine if Mtd was associated with cell death in the molar tissue, we assessed the protein

expression of Mtd in conjunction with a number of molecules involved in apoptosis. Cleaved

caspase 8, cleaved caspase 3 and cleaved Parp1 are protein products commonly used to indicate

the presence of apoptosis, whereas Mcl-1 is an anti-apoptotic Bcl-2 family member known to

play an important role in trophoblast survival. Contrary to what we had anticipated, Mtd was the

only pro-apoptotic molecule tested that was found to be increased in the molar samples (Figure

5.4b-d). In contrast cleaved-Parp1, cleaved caspase-3, and cleaved caspase-8 were decreased or

unchanged in the molar placentae compared to either their co-existing twin or the normal twin

control sets (Figure 5.4a). In addition, no differences were observed for Mcl-1 between the

molar samples and the co-existing twin control (Figure 5.4b). We next assessed the transcript

level of Mtd-L to determine if the elevated levels of Mtd were due to increased gene

transcription. Mtd-L mRNA expression was increased in case 1 relative to its co-twin but similar

expression was observed between the molar and the control co-twin placentae for case 2 (Figure

5.4c).

Since Mtd is associated with cell death in the placenta, we performed co-localization studies of

Mtd and TUNEL to determine if Mtd was co-expressed in cells with fragmented DNA in the

molar pathology. As expected, Mtd expression was associated with TUNEL positive cells in the

molar placentae (Figure 5.4d upper panels). The placentae from the co-twin and normal twin

controls also displayed co-expression of Mtd and TUNEL however fewer TUNEL positive cells

were found in these samples (Figure 5.4d lower panels, data not shown). Interestingly cells

122

Figure 5-4 Expression of apoptotic molecules in molar twins and control twins

a: Representative Western blot of apoptotic molecules: cleaved-Parp1, cleaved caspase-3, cleaved caspase-8, and b:

Mtd and anti apoptotic Mcl-1 in placental lysates from molar twin cases, and normal age matched twin controls.

Actin expression was used as an internal control. c: qRT-PCR analysis of MtdL mRNA from total placental tissue

lysates from molar twins (Mole, filled bar, n=2; co-existing twin, unfilled bar, n=2), and AMC twins (Con:A and

Con:B, hatched bars, n=7 twin sets). Data were normalized against expression of 18S ribosomal RNA using the well

established 2-CT

formula; fold changes are relative to AMC singleton pregnancy internal control group. Statistical

significance was assessed between control twins. P<0.05 Mann Whitney U test d: Spatial localization of Mtd

(green) with TUNEL (red) in placental sections from molar case 1 (left panels), and molar case 2 (right panels).

Arrow: cells positive for TUNEL reactivity

123

that were TUNEL positive displayed predominantly cytoplasmic Mtd staining whereas Mtd was

primarily nuclear in cells that were TUNEL negative (Figure 5.4d).

5.3.4 Mtd localizes to the nuclei of proliferative trophoblast cells in molar

pathology

To determine if the pattern of Mtd expression seen in molar pregnancy was reflective of the role

of Mtd in proliferation, we performed dual labeled immunofluorescence analysis with antibodies

against Mtd and Ki67 (Figure 5.4 and 5.5). Similar to our previous reports, co-expression of

Mtd and Ki67 was localized to the cytotrophoblast cells in control sections (Ray et al., 2009)

(Figure 5.5b-d). An increased number of trophoblast cells in the molar tissue co-expressed Mtd

and Ki67 compared to the co-existing twin or normal twin control placentae (Figure 5.5a). In the

molar placentae Mtd-Ki67 co-localization could be detected underlying areas of excessive

syncytial accumulation, (Figure 5.6-i) throughout hyperproliferative sprouts (Figure5.6-ii.iii),

and in thinned trophoblastic areas (Figure 5.6-iv).

5.3.5 Mtd is associated with increased cyclin E1 in villous trophoblast cells of

molar placentae

Since the molar samples exhibited excessive proliferation, and displayed co-expression of Mtd

with Ki67 positive trophoblast cells, we tested the level of cyclin E1, a G1 phase cell cycle

regulator that we have previously reported to be associated with Mtd expression in the placenta

(Ray et al., 2009). In both molar sets, the samples from the molar tissue displayed higher levels

of cyclin E1 protein and mRNA expression compared to the placentae from their co-existing

twin controls (Figure 5.7a,b respectively). To test whether Mtd was co-localized with cyclin E1

in the molar pathology we performed dual labeled immunofluorescence with antibodies against

Mtd and cyclin E1 (Figure 5.7c). Similar to that seen in preeclampsia Mtd was expressed in

cyclin E1 positive trophoblast cells in the molar placentae (Figure 5.7c upper panels)(Ray et

al., 2009). Co-localization of Mtd and cyclin E1 was also observed occasionally in the placentae

of the co-existing twins and the normal twin controls, but to a lower extent (data not shown). In

addition cyclin E1 could be detected around the vasculature in the control sections (Figure 5.7c

lower panels), a characteristic absent in the molar pathology.

124

Figure 5-5 Co-localization of Mtd with Ki67 expression in mole and twin placentae

Co-localization of Mtd (green) with Ki67 (red) in placental sections from molar case 2 (top left), and in the placenta

of its co-existing twin (top right), and in the placentae of a representative set of age matched control twins (bottom

panels). Right hand panels: high magnification of the boxed areas. Arrow: cytotrophoblast cells expressing both Mtd

and Ki67

125

Figure 5-6 Mtd is expressed in proliferative trophoblast cells associated with various molar characteristics

Co-localization of Mtd (green) with Ki67 (red) in (i) areas of excessive syncytial accumulation, (ii-iii)

hyperproliferative sprouts, and (iv) in thinned areas of trophoblast. Middle and right hand panels: high magnification

of the boxed areas. Arrows: trophoblast cells expressing both Mtd and Ki67

126

Figure 5-7 Cyclin E1 is overexpressed in molar placentae compared to twin controls

a: Representative Western blot of cyclin E1 in placental lysates from molar twin cases, and normal age matched

twin controls. Actin expression was used as an internal control. b: qRT-PCR analysis of cyclin E1 mRNA from total

placental tissue lysates from molar twin sets (Mole, filled bars, n=2; co-existing twin, unfilled bars, n=2), and AMC

twins (Con:A and Con:B, hatched bars, n=7 twin sets). Data were normalized against expression of 18S ribosomal

RNA using the well established 2-CT

formula; fold changes are relative to group of age matched singleton

pregnancy. Statistical significance was assessed between control twins. P<0.05 Mann Whitney U test c: Spatial

localization of Mtd (green) with cyclin E1 (red) in placental sections from molar case 2 (top panels), and in the

placenta of its co-existing twin (bottom panels). Right hand panels: high magnification of the boxed areas. Arrows:

cytotrophoblast cells expressing both Mtd and cyclin E1; Arrowheads: cyclin E1 expression in endothelium.

127

5.3.6 Molar placentae exhibit decreased levels of cell cycle inhibitor p27

As shown in chapter 3, the CDK2 inhibitor, p27 is highly expressed in association with high

levels of cyclin E1 in placentae from severe early onset preeclamptic pregnancies. We therefore

investigated the expression of p27 in the molar pathology. In contrast to early onset severe

preeclampsia, lower levels of p27 protein were observed in molar placentae compared to their

co-existing twin or the normal twin controls (Figure 5.8a). Co-localization of p27 with cyclin E1

also demonstrated p27 expression to be decreased in the cyclin E1 positive cells in the molar

tissue in comparison to the controls (Figure 5.8b).

5.3.7 Molar placentae exhibit altered expression of molecules involved in

regulating the G1 phase of the cell cycle

Next we assessed the expression of D type cyclins in the placenta as well as the CDK inhibitor

p21, to determine whether the increased proliferation in the mole was facilitated by cell cycle

events in the G1 phase additional to the increase in cyclin E1. Both cyclin D1 and cyclin D3

were decreased at both protein and mRNA levels in the molar pathologies relative to the

placentae of the co-existing twin from case 2 and the control twin sets (Figure 5.9a,b).

Interestingly, expression of cyclin D1 and cyclin D3 in the molar placentae was predominantly

localized to the stroma (Figure 5.9c,d), whereas in control sections cyclin D1 was expressed

primarily in the cytoplasm of the syncytiotrophoblast (Figure 5.9c) and cyclin D3 was expressed

in the nuclear compartment of cells of the trophoblastic layer (Figure 5.9d). In contrast to the

expression of the D type cyclins or p27, p21 expression was increased in the molar samples

compared to the placentae of their co-existing twins (Figure 5.9a). Localization of p21 however

was restricted to the nuclei of trophoblast cells in both molar and control placental sections

(Figure 5.9d). Of note, the overall level of protein expression for cyclin D1, cyclin D3 and p21,

was low in case 1 (Figure 5.9a). No significant differences in cyclin D1, D3 or p21 expression

were seen within or between normal twin placental controls (Figure 5.9a).

5.4 Discussion

The placenta shares similar features to that of a controlled cancer, with highly proliferative and

invasive properties. However, unlike a cancer it is intrinsically regulated to regress and evade

pathological progression. In molar pregnancy, this regulation is altered, causing uncontrolled cell

128

Figure 5-8 p27 expression in molar placentae and twin controls

a: Representative Western blot of p27 in placental lysates from molar twin cases, and normal age matched twin

controls. Actin expression was used as an internal control. b: Spatial localization of p27 (green) with cyclin E1 (red)

in placental sections from molar case 2 (top panels), and in the placenta of its co-existing twin (bottom panels).

Right hand panels: high magnification of the boxed areas. Arrow: cell expressing cyclin E1, Arrowhead:

cytotrophoblast cells expressing both p27 and cyclin E1.

129

Figure 5-9 Expression of G1 phase cell cycle regulators in the molar and control twins

a: Representative Western blot of: cyclin D1, cyclin D3, and p21 in placental lysates from molar twin cases (M,Tw),

and AMC twins (A,B). Actin was used as an internal control. b: qRT-PCR analysis of cyclin D1 and cyclin D3 in

total placental tissue lysates from molar twin cases (Mole, filled bars, n=2; co-existing twin, unfilled bars, n=2), and

AMC twins (Con:A and Con:B, hatched bars, n=7 twin sets). Data were normalized against 18S using the well

established 2-CT

formula; fold changes are relative to singleton AMC group. Statistical significance was assessed

between control twins. P<0.05 Mann Whitney U test c,d: Spatial localization of c: cyclin D1 (red) or d: p21 (green)

with cyclin D3 (red), in placental sections from molar case 2 (top left), and in the placenta of its co-existing twin

(top right), and in a set of age matched control twins (bottom panels). Right hand panels: high magnification of the

boxed areas. c: Arrow: syncytiotrophoblast; Arrowhead: cytotrophoblast cells negative for cyclin D1. d: Arrow:

trophoblast cells positive for cyclin D3; arrowhead: trophoblast cells positive for p21.

130

behavior, and in many cases leads to malignant transformation. This can result in persistent

trophoblastic disease and in the most severe cases choriocarcinoma. Herein, we examined

molecules known to play key roles in regulating normal trophoblast cell fate, to gain further

insight in to the pathogenesis of molar pregnancy. Specifically we discovered 1) that Mtd, a pro-

apoptotic Bcl-2 family member, is increased in molar placentae, 2) that the expression of Mtd in

molar tissue is associated with both hyperproliferative and apoptotic characteristics of placental

cells, in particular that the hyperproliferation of the trophoblast cells are associated with an

imbalance between low p27 and high cyclin E1 expression, and 3) that decreased expression of

cyclin D1 and cyclin D3 are associated with increased p21 expression in the trophoblast layers of

molar tissue.

This study reported on two unique cases of molar pregnancy consisting of a complete

hydatidiform mole and a co-existing fetus. These molar twin cases are extremely rare with a

frequency reported at 1 in 22,000 to 1 in 100,000 pregnancies (Vaisbuch et al., 2005). A number

of studies have reported on singleton molar pregnancy from the early first trimester, however,

due to the immediate evacuation of these placentae once identified, molar tissue at later

gestations have not been studied at the molecular level. Interestingly, the second trimester molar

cases presented in this study displayed classic morphological signs of molar villi, including

vessel swelling, circumferential hyperplasia and cistern formation. Molar placentae typically

present with few or no villous blood vessels, and those present are usually collapsed and empty

(Kim et al., 2006). The two cases presented here also had no obvious vessel architecture

indicative of progressed vessel deterioration. In addition it is characteristic for molar placentae to

have apoptotic debris in the stromal area, as seen in the two cases presented. Pathological

description of these two cases suggests that second trimester molar twin cases are grossly

equivalent to first trimester moles that develop as a singleton pregnancy.

Importantly, these rare, molar-twin cases provide a unique model to study the molecular

pathophysiology of molar development, as each molar case developed adjacent to a genetically

normal twin placenta, used as a baseline control in the analysis of the data. Our findings are

therefore highly reflective of the innate genetic and molecular differences between molar tissue

development and placental development, represented by the co-existing twin placenta. However,

due to the complications associated with the molar twin pregnancies, the molecular findings were

also compared to normal age-matched twin controls. Interestingly the first case of molar twins

131

displayed low levels of all the molecules examined, with the exception of TUNEL reactivity.

This finding suggests that the pregnancy in case 1 was severely compromised, possibly due to

the bacterial infection. Nevertheless, when the molar tissue in case 1 was compared to its co-twin

placenta, the trends were found to be similar to that of case 2. In case 2 the genetically normal

co-placenta displayed slight elevations in the apoptotic markers cleaved Parp1 and cleaved

caspase 8, as well as elevated levels of cyclin E1 and p27 in comparison to the control twin sets.

The increase in apoptotic markers likely reflects the clinical finding that the co-twin in case 2

was IUGR and presented with preeclampsia, two pathologies associated with increased placental

apoptosis. Likewise the increase in cyclin E1 and p27 may also reflect the incidence of early

onset preeclampsia, as seen in chapter 3.

Importantly, molar tissue provides a unique model to study the mechanisms and molecules that

regulate both cell proliferation and cell death. We have previously reported on Mtd, a pro-

apoptotic member of the Bcl-2 family, to be highly expressed in the early stages of normal

placental development and in severe early onset preeclampsia, where it was found to regulate

apoptotic cell death as well as cell proliferation (Soleymanlou et al., 2005b;Ray et al., 2009).

The high level of Mtd expression found in molar tissue in the current report was therefore not

surprising, as molar pregnancy represents an exemplary pathological state of excessive

trophoblast proliferation and placental apoptosis.

Previous reports have described apoptosis occurring in both the cytotrophoblast layer and in the

villous stroma in first trimester molar pregnancy (Halperin et al., 2000;Kim et al., 2006).

Similarly, in this study Mtd was found to co-localize to TUNEL positive cells in both the

cytotrophoblast and stromal region supporting the idea that Mtd plays a role in apoptosis in this

pathology. In the stroma of molar tissue, cell death has been associated with the destruction of

the primitive vascular network, resulting in the fluid accumulation and villous swelling (Kim et

al., 2006), as well as the classic acellular stromal characteristic. Whereas apoptosis in the

trophoblast layers is associated with the exacerbated proliferation, similar to that seen in

preeclampsia. It is therefore possible that Mtd may contribute to both the acellular and avascular

stromal feature of molar development, as well as the excessive trophoblastic shedding and

subsequently high incidence of preeclampsia in this pathology.

132

In addition, a number of studies have linked the level of apoptosis in molar pregnancy with the

potential for malignant transformation, where high apoptotic rates are suggested to be indicative

of a favorable prognosis (Qiao et al., 1998;Wong et al., 1999;Chiu et al., 2001). These findings

have been based primarily on TUNEL reactivity and the detection of caspase-cleaved

cytokeratin, although additional studies have suggested that the expression of Bcl-2 family

members may also provide further insight as to the aggressiveness of this pathology. It has been

theorized that the Bcl-2/Bax ratio may be an important indicator of overall cell death in the mole

(Qiao et al., 1998). However, a more recent study reported that Mcl-1 expression displayed a

more significant correlation to the prognosis of gestational trophoblastic disease (GTD)

compared to either Bcl-2 or Bax (Qiao et al., 1998;Fong et al., 2005). High levels of Mcl-1 were

found to be indicative of a poor prognosis, likely due to the anti-apoptotic effect of the protein.

Importantly Mcl-1 is the primary binding partner of Mtd. The interaction of Mcl-1 with Mtd

prevents the apoptotic effect of Mtd and has been shown to lead to the decreased incidence of

cell death (Soleymanlou et al., 2007). Since molar tissue expresses high levels of Mtd, the

Mtd/Mcl-1 rheostat may be of significance to molar development and progression. Interestingly,

in the two cases presented we found elevated Mtd expression and no difference in Mcl-1 levels

compared to control; an Mtd/Mcl-1 ratio that in theory would be favorable for high levels of cell

death and thus molar regression. This is supported by the high incidence of apoptosis seen in our

cases. A follow up study of these patients would be required to confirm that molar regression

resulted in these cases following molar evacuation, however in case 2 a full hysterectomy was

performed preventing further follow up analysis.

In general, apoptosis can result from activation of the extrinsic or intrinsic apoptotic pathways

involving the caspase cascade. Activation of caspase initiators (casp-2, 8, 9,10) leads to cleavage

and activation of the executioner caspases such as caspase 3 and 7 directly or through

mitochondrial depolarization (intrinsic pathway, involving members of the Bcl-2 family).

Activation of downstream executioner caspases results in cellular changes, such as cytoskeletal

rearrangement, due to cleavage of a variety of cellular proteins including Parp1. Interestingly,

although the clinical cases presented demonstrated an increased Mtd/Mcl-1 ratio and increased

TUNEL positivity, the majority of the apoptotic molecules examined (cleaved caspase-8, cleaved

caspase-3, cleaved Parp1) displayed lower or equivalent levels of expression compared to

controls. This suggests that the majority of the molecules tested may not be significant

133

contributors to apoptosis in the mole. Similarly Wong 1999 and Qiau 1998 found that Bax had

no correlation with the apoptotic index in molar tissue and Fong et at 2006 found no difference

in expression of either caspase-8 or caspase-10 between molar tissue and placentae from normal

pregnancy. It is therefore possible that apoptosis in molar tissue is carried out through related

caspases such as initiator caspases 2 or 9 and executioner caspase 7, not examined in this study.

Further examination of these molecules in association with Bcl-2 family members such as Mtd

and Mcl-1in molar tissue is therefore warranted.

p53 is another pro-apoptotic molecule involved in both cell death and cell proliferation and,

similar to Mtd, it is elevated in the trophoblast cells of molar tissue (Qiao et al., 1998;Halperin et

al., 2000). Interestingly, mutation of the p53 gene has been associated with its abundance in a

number of cancers, however, in molar tissue it is the wild type form of p53 that is overexpressed

(Cheung et al., 1994;Halperin et al., 2000). Importantly, p53 has been shown to induce Mtd

expression, and it may therefore contribute to the elevated levels of Mtd seen in this pathology

(Yakovlev et al., 2004). In addition, studies have shown that p53 may be important in the

neoplastic nature of the mole (Qiao et al., 1998). It would therefore be interesting to compare the

level of Mtd in molar samples that regress to those that persist, to examine whether Mtd has a

similar effect. This area of study warrants further investigation.

In the current report Mtd co-localized to both proliferative (Ki67 and cyclin E1 positive) as well

as apoptotic (TUNEL positive) cells in the two molar cases, suggesting that in molar tissue Mtd

functions in promoting cell cycle progression in addition to cell death, similar to its function in

normal placentation and in preeclampsia. Furthermore, in the first data chapter we found Mtd to

influence the cell cycle through its positive impact on cyclin E1 (Ray et al., 2009). Our finding

that Mtd co-localized with Ki67 and cyclin E1 in trophoblast cells in molar tissue, therefore

suggests that Mtd may contribute to the hyperproliferative nature of the molar pathology by

promoting the expression of cyclin E1. This was supported by the finding that the two molar

cases displayed increased levels of cyclin E1 and Ki67 compared to the co-existing twin or

control twin sets. Similarly, Ki67 and cyclin E1 have also been found to be increased in first

trimester molar placentae compared to normal placental development of the same gestation

(Fukunaga, 2004). Olvera et al 2001 had comparable findings for cyclin E1 and Ki67 and in

addition reported increased levels of Cdk2 and E2F1 in the trophoblastic layers of molar tissue

(Olvera et al., 2001). Activation of Cdk2 by cyclin E1 leads to the release of E2F1 from its

134

inhibitor, Rb, and the subsequent transcription of E2F1 regulated genes. Interestingly, E2F1

regulates transcription of both cyclin E1 and Mtd, and this may explain the high levels of these

molecules observed in our second trimester molar cases.

In addition to Rb, the Cyclin E-CDK2 complex also phosphorylates the CDK inhibitor p27 at the

Thr187 site leading to its destruction in the nucleus (Vlach et al., 1997). This may explain the

low levels of p27 observed in the molar tissue. Furthermore, the low level of p27 maintained in

the trophoblast cells is likely insufficient to fully inhibit cell cycle progression. Interestingly, the

low level of p27 seen in the molar pathology is in contrast to the elevated levels of p27 seen in

severe early onset preeclampsia in the previous chapter. This observation indicates that although

the two placental pathologies share similar characteristics, there are likely key differences in

upstream events that underlie the two pathologies. Nevertheless, in both preeclamptic and molar

placentation the inhibitory effect of p27 appears to be hampered thus, contributing to the

proliferative nature the disorders.

Although low levels of p27 were observed in the two molar cases presented, we did see an

increase in expression of the related CDK inhibitor p21 in the trophoblastic layers. p21 can

hinder cell cycle progression by inhibiting the function of the cyclin E1-CDK2 complex, or it can

function in the assembly of cyclin D-CDK complexes and aid in promoting cell cycle

progression (Labaer et al., 1997;Cheng et al., 1999). Interestingly, in the control samples p21

was occasionally co-localized with cyclin D3. It is therefore possible that p21 functions in the

assembly of cyclin D3-CDK complexes in normal placental development. In contrast, no cyclin

D1 or cyclin D3 was observed in the trophoblastic layers in the molar tissue. This suggests that

the high levels of p21 may function to inhibit, rather than promote, cell cycle progression in the

trophoblast cells of molar tissue. Furthermore, p21 expression was primarily restricted to the

more distal trophoblast cells in the molar pathology, consistent with a role in differentiation or

quiescence. Future studies are warranted to determine if p21 functions in the differentiation of

the trophoblast cells in molar tissue. If so it could serve as a good prognostic indicator against

malignant transformation in molar cases. However, Cheung et al 1998 have previously reported

that there was no significant difference in the expression of p21 between cases that persisted and

those that regressed (Cheung et al., 1998).

135

In the current study, we found that the molar tissue expressed very low levels of cyclin D1 and

cyclin D3. In addition, the small amount of protein expressed in the mole was restricted

primarily to the stromal tissue, whereas in normal placental tissue both cyclin D1 and cyclin D3

were expressed by the trophoblastic layers. Expression of nuclear cyclin D1 in the stromal cells

of the molar tissue is similar to the pattern of expression seen for cyclin D1 in early first

trimester (data not shown) suggesting that the stromal tissue in molar placentae may remain in an

undifferentiated state. Cytoplasmic cyclin D1, such as that seen in the syncytium of the control

sections, has been linked to cell differentiation of murine embryonic stem cells (Bryja et al.,

2008). Therefore the lack of cyclin D1 in the trophoblast of the mole may also reflect a poorly

differentiated state of the molar tissue. Alternatively the low level of cyclin D1 expression may

be due to increased expression of E2F-1 which has been shown to inhibit activation of the cyclin

D1 promoter (Watanabe et al., 1998).

Interestingly, reports on murine models suggest that D type cyclins are not crucial for placental

development, as cyclin D triple knockout mice were found to have no defects in placentation

(Kozar et al., 2004). In addition, certain cell types, including some embryonic tissues, have been

shown to proliferate efficiently in the absence of D type cyclins (Kozar et al., 2004). D type

cyclins are normally controlled by mitogenic stimuli and function to link cues from the extra

cellular environment with control over cell proliferation. The fact that trophoblast cells in the

mole proliferate independent of cyclin D expression suggests that molar tissue may be unaffected

by external cues. In addition these finding also suggest that molar tissue could be used to study

the difference in cyclin D dependent and cyclin D independent cell cycle machinery.

Many studies have focused on uncovering traits that will distinguish whether a molar case will

become persistent or if it will regress. Although the presence, or absence, of cell cycle regulating

molecules have been associated with oncogenic transformation, (Sgambato et al.,

2000;Slingerland et al., 2000;Kozar et al., 2004) studies have yet to find a cell cycle regulating

molecule that has prognostic value in the mole. There is however evidence that the level of

apoptosis may be a reliable diagnostic tool. It is therefore important to learn more about the

proliferative and apoptotic activity in the pathophysiology of molar pregnancy so that diagnostic

methods and clinical treatments can be identified.

136

6 Summary and Future Directions

Proper formation of the placenta is essential for fetal development and a successful healthy

pregnancy. In humans, this process involves a balance in cellular proliferation, differentiation

and death of the trophoblast lineage, the cells forming the placenta. Consequently, abnormality in

the regulation of trophoblast cell fate may lead to inadequate placental function and subsequent

pregnancy related complications.

Preeclampsia is a life threatening disorder of pregnancy, common in the obstetric field. It has

been well established that the cause of preeclampsia stems from the placenta, however to date,

no reliable prognostic or diagnostic tools for preeclampsia exist. Moreover, the current standard

of treatment for preeclampsia remains delivery and removal of the placenta. Prior research has

established that the maternal symptoms of preeclampsia arise from inflammation and endothelial

dysfunction, occurring from a reaction of the maternal endothelium to placental-derived factors.

In cases of severe early onset preeclampsia this is believed to occur due to excessive deportation

of placental debris into the maternal circulation. In contrast, late onset preeclampsia appears to

result from a sensitivity of the maternal endothelium to normal levels of placental debris, and is

thus a maternally derived disorder. In either case, women with preeclampsia may require prompt

delivery of the fetus, causing potential developmental complications to the newborn.

Complete molar pregnancy is yet another placental disorder. In these cases, placental tissue

develops in the absence of a fetus, and presents a high risk of developing preeclampsia and

choriocarcinoma requiring chemotherapy. Although separate pathologies, both preeclampsia and

molar pregnancy share similar characteristics including an immature proliferative trophoblast

phenotype accompanied by excessive cell death and increased trophoblast turnover. Deciphering

the mechanisms that regulate cell proliferation and cell death in the placenta will therefore allow

us to gain insight into the pathogenesis of both these pregnancy-related disorders.

The objective of the current dissertation was to further our knowledge on the systems governing

the balance between cell death and cell proliferation in the placenta and to identify alterations in

the expression of cell fate regulatory molecules that may contribute to placental pathologies.

Work presented in this thesis provides novel molecular insights into the regulation of trophoblast

cell proliferation by Mtd and cell cycle regulating molecules in normal placental development as

137

well as in preeclampsia and molar pregnancy. Importantly the data presented in this dissertation

also describe molecular differences between clinically related but distinct placental pathologies,

further advancing the potential for future diagnostic, prognostic and treatment strategies.

The first data chapter of the dissertation (Chapter 3) examined the role of Mtd, a pro-apoptotic

molecule, in normal placental development. This line of investigation identified a novel role for

Mtd in the regulation of trophoblast cell cycle in placentation. The dual roles of Mtd were

associated with changes in its subcellular localization, with Mtd localizing to the mitochondria

during apoptosis, and to the nucleus in proliferative cells. Although the current dissertation

revealed insight as to how cell death and proliferation in the trophoblast may be interconnected

by Mtd, we recognize that these processes are extremely complex and involve a multitude of

players. Recently a number of classical apoptotic and cell cycle regulating molecules have been

identified as having multiple roles related to cell fate. Caspases (cysteine-aspartate proteases) are

among this group. Although caspases are classically known for their prominent role in apoptosis,

recent studies have revealed a variety of non-apoptotic functions of these cysteine proteases,

including roles in cell differentiation, cell motility, and importantly cell cycle regulation

(Feinstein-Rotkopf and Arama, 2009). In future studies, it would be of value to determine if

caspases, in conjunction with Mtd, are involved in the regulation of trophoblast cell cycle in the

human placenta.

Studies presented in this dissertation also revealed that the Mtd-L isoform had a positive effect

on cyclin E1 expression and promoted G1-S phase transition. However the means by which Mtd

impacts cyclin E1 expression remains unknown. As such, it would be of significance to

determine the mechanism linking Mtd to cyclin E1 expression. The Caniggia lab has previously

shown that active caspase-3 is a downstream effecter of Mtd function during apoptosis

(Soleymanlou et al., 2005b). Importantly, a number of molecules involved in cell cycle

regulation, including cell cycle inhibitors p21 and p27, are reported to be cleaved by caspases

(Eymin et al., 1999;Woo et al., 2003). To date, caspase-3 has been found to promote cell cycle

progression in lymphoid cells, forebrain cells and keratinocytes. In lymphoid cells this has been

attributed to caspase cleavage of cell cycle inhibitor p27. In contrast, caspase-3 has been found to

inhibit proliferation of B cells through a mechanism involving cleavage of p21 (Waga et al.,

1994;Woo et al., 2003). The data presented in this dissertation suggests that p27 is a prominent

antagonist to cyclin E1 in the placenta. Therefore it would be of interest to determine if caspase

138

mediated cleavage and subsequent inactivation of p27 in trophoblast cells may be the link

between Mtd and cyclin E1 expression. A putative model is depicted in Figure 6.1.

In the second and third data chapters (Chapters 4 and 5), an association between Mtd and

increased levels of cyclin E1 was discovered in cytotrophoblast cells from placentae of early

onset severe preeclampsia and molar pregnancy. These findings demonstrate that Mtd plays an

intricate role in maintaining the natural homeostasis in trophoblast cell fate, and it suggests that

the increased expression of Mtd previously seen in preeclampsia (Soleymanlou 2005) likely

contributes to both the hyperproliferative and apoptotic nature of the placenta typical of this

disorder.

Studies presented in Chapter 4 of the dissertation also addressed the cause of the

hyperproliferative state of trophoblast cells in preeclampsia at the level of the cell cycle. First, in

order to understand the potential consequences of elevated cyclin E1 expression in placental

disease, its role and regulation in normal placentation were assessed. This series of studies

indicated that cyclin E1 was maximally expressed in the early stages of placentation where its

expression was inversely correlated with that of the CDK inhibitor p27. We also determined that

a low oxygen environment in conjunction with high levels of TGF3, such as that previously

found in early first trimester and in the preeclamptic pathology, can lead to increased cyclin E1

and p27 protein expression. In light of this data we hypothesized that placental pathologies, such

as preeclampsia, would display increased Mtd and cyclin E1 and would exhibit low levels of p27

resembling the molecular profile of an immature early first trimester placenta. In contrast to our

hypothesis, we discovered an increased level of p27 expression in severe early onset

preeclampsia. However, p27 in this pathology was localized primarily to the cytoplasm.

Moreover, this led to the identification of elevated levels of p27 phosphorylation at the Ser10 site

in severe early onset preeclampsia. We speculated that this post-translational modification led to

the observed increase in cytoplasmic localization of p27, preventing its inhibition on cyclin E1-

mediated CDK2 activation. We further hypothesized that this may contribute to the increased

trophoblast turnover and syncytial shedding seen in preeclampsia. We did not however examine

the mechanism by which p27 was phosphorylated in the preeclamptic pathology. Excessive

trophoblast deportation is one of the primary factors leading to endothelial dysfunction in severe

early onset preeclampsia. Thus, identifying the upstream factors leading to the increased

139

Figure 6-1 Putative model of the mechanism linking Mtd to cyclin E1 expression

Mtd initiates mitochondrial depolarization leading to caspase-3 cleavage and activation. Activated caspase-3 cleaves

the CDK inhibitors p21 and p27 alleviating this inhibition from the cyclin E-CDK2 complex. The activated cyclin E

CDK2 complex then phosphorylates Rb which in turn dissociates from the E2F transcription factor. Once free, E2F

activates the transcription of multiple genes, including Mtd and cyclin E1, thereby creating a positive feedback loop.

140

trophoblast turnover may aid to the discovery of potential treatments to help alleviate the clinical

symptoms of preeclampsia and thus the overall severity of the disease.

Ser10 phosphorylation of p27 has been reported to occur at the G1 phase downstream of KIS,

PKB/Akt, and ERK signaling (Besson et al., 2008) (Figure 6.2). Phosphorylation at this site

during the G1 phase promotes nuclear export by providing a binding site for CRM1/Exportin.

Alternatively, in G0, Ser10 phosphorylation results from activation of Mirk/Dyrk kinase

therefore conferring p27 stability in the nucleus (Deng et al., 2004). Since our observations in

preeclampsia are consistent with an increase in G1 phase activity, it would be relevant to

examine the pathways that have been reported to phosphorylate p27 during the G1 phase.

hKIS (human kinase interacting stathmin) was discovered in 2002 to be the major kinase that

phosphorylates Ser10 in p27 (Boehm et al., 2002). In addition, it was shown to promote cell

cycle progression in mouse embryonic fibroblast cells as well as human leukemia cells

(Nakamura et al., 2008) by stabilizing p27 in the cytoplasm during the G1 phase. Interestingly,

Boehm et al also found that hKIS was highly expressed in the placenta; supporting our

hypothesis that p27 may be phosphorylated by hKIS during placentation. To our knowledge

there are no reports on hKIS activity in the placenta or on the expression of hKIS in

preeclampsia. It would therefore be informative to determine if hKIS phosphorylates p27 in

trophoblast cells and to examine whether hKIS levels are altered in severe early onset

preeclampsia.

A second pathway described to phosphorylate p27 Ser10 in the G1 phase is the PKB/Akt

pathway (Fujita et al., 2002;Besson et al., 2004). However, it is unlikely that the increased p27

phosphorylation seen in our preeclamptic samples results from Akt phosphorylation, as the Akt

pathway is reported as being either unaltered or diminished in this pathology (Orcy et al.,

2008;Chiang et al., 2009). Alternatively, p27 could be phosphorylated at Ser10 by ERK

(extracellular signal-regulated kinase) downstream of the MAPK (mitogen activated protein

kinase) pathway (Ishida et al., 2000). This is supported by Shin et al who have reported ERK to

be increasingly active in severe preeclampsia (Shin et al., 2009).

Our studies also revealed that severe early onset preeclampsia is associated with cytoplasmic

localization of p27, and we hypothesize that p27 is stabilized in the cytoplasm in this pathology.

Stabilization of p27 in the cytoplasm has been reported to occur through its interaction with

141

Figure 6-2 Phosphorylation of p27 at the ser10 site.

Ser10 phosphorylation of p27 occurs in the G0 phase by the Mirk/Dyrk kinase and in the G1 phase downstream of

KIS, PKB/Akt, and ERK signaling. Phosphorylation at this site during the G1 phase promotes nuclear export by

providing a binding site for CRM1/Exportin. In G0, ser10 phosphorylation confers p27 stability in the nucleus by

binding to Mirk/Dyrk.

142

cyclin D-CDK complexes or by further phosphorylation at its Thr-157 or Thr-198 sites, leading

to its interaction with 14-3-3 (Besson et al., 2008;Vervoorts et al., 2008) (Figure 6.3).

Determining the mechanism that facilitates p27 stability in the cytoplasm in preeclampsia may

be of interest, as therapeutic treatment targeted at disrupting the stabilizing factor may allow p27

to reenter the nucleus and prevent trophoblast hyperproliferation.

The comparison of cyclin E1 and p27 expression, between preeclampsia-related pathologies

including late onset preeclampsia and IUGR was another important study presented in the

dissertation. Here we discovered that the high levels of cyclin E1 and cytoplasmic p27 were

unique to the severe early onset form of preeclampsia, whereas neither feature was associated

with either late onset preeclampsia or IUGR. These findings indicate that cyclin E1, in addition

to altered p27 regulation, aids in the hyperproliferative state of severe early onset PE and that

this may be potentiated by the low pO2 and elevated levels of TGF in the preeclamptic

environment. These molecular data also support the theory that severe early onset preeclampsia

is a distinct pathology from late onset preeclampsia. A further understanding of the differences

between placental-related pathologies will provide a basis with which to improve diagnostic

methods in the future and aid in tailoring more direct and effective approaches to treatment.

In the final chapter two unique cases of second trimester twin pregnancies, including a complete

mole and co-existing twin, were examined. Through these studies Mtd was identified as being

highly expressed in the molar pathology, localizing to both apoptotic cells present in the

trophoblast and stromal regions, as well as to the hyperproliferative areas of trophoblast cells. In

addition, Mtd expression was abundant in trophoblast cells positive for cyclin E1, similar to what

we have shown in severe early onset preeclamptic cases. In contrast to the elevated levels of p27

seen in severe early onset preeclampsia, molar tissue displayed decreased levels of p27. We

therefore propose that the high abundance of Mtd and cyclin E1 in conjunction with the low level

of p27, may contribute to the hyperproliferative nature of the disorder. Low levels of p27 are

commonly associated with various cancers, resulting primarily from increased proteolytic

degradation rather than allelic loss (Sherr et al., 2004). Hence, it would be informative to

determine if the low level of p27 associated with molar development results from altered p27

transcription or from increased proteolytic degradation of the protein.

143

Figure 6-3 Phosphorylation of p27 and the effect on protein localization and stability

The cyclin E1-CDK2 complex phosphorylates p27 at the Thr187 residue revealing a binding site for Skp2-SCF.

Ser10 phosphorylation of p27 in the G0 phase by the Mirk/Dyrk kinase confers p27 stability in the nucleus by

binding to Mirk/Dyrk. Phosphorylation at ser 10 by KIS, PKB/Akt, or ERK during the G1 phase promotes nuclear

export by providing a binding site for CRM1/Exportin. P27 is stabilized in the cytoplasm by interaction with cyclin

D-CDK complexes or by further phosphorylation at Thr157 or Thr198 leading to its interaction with 14-3-3.

Interaction of p27 with RhoA is involved in cell migration.

144

Both severe early onset preeclampsia and molar pregnancy are hyperproliferative placental

pathologies that display elevated levels of trophoblast apoptosis. Importantly, in this dissertation

we discovered that placentae from both disorders displayed similarities in their molecular profile,

including increased Mtd and cyclin E1 expression. However, our data also highlighted the

differences that occur in cell cycle regulation between the pathologies, namely the differences in

p27 regulation. A putative model comparing Mtd, cyclin E1 and p27 expression in molar

pregnancy and severe early onset preeclampsia is summarized in (Figure 6-4). This underscores

the importance of investigating cell cycle regulation in independent placental pathologies, as

potential treatments strategies may have different effects in each case.

Current clinical practice following evacuation of complete molar tissue is to monitor the human

chorionic gonadotropin (HCG) levels in the women in order to determine if their molar

pathology fully regressed. Remaining HCG levels indicate persisting tissue, and only then can

the physician plan additional treatment, often including chemotherapy. This current method of

practice is time consuming for both the patient and the physician and most importantly can result

in a lengthy period of uncertainty and anguish for the already devastated woman. It is therefore

imperative that research continues toward discovery of prognostic indicators of molar pregnancy

outcomes. Previous studies suggest that cell death regulating molecules may be the link to the

behavior of molar tissue. Both TUNEL positivity and Mcl-1 have been hypothesized to be

potential indicators of molar regression and persistence (respectively) (Qiao et al., 1998;Wong et

al., 1999;Chiu et al., 2001). Importantly, Mtd is a pro-apoptotic molecule that interacts with Mcl-

1, and since we found Mtd to be increased in molar pathology, it would be worth studying the

relationship between Mtd and molar progression in cases that regress compared to those that

persist.

Further investigation into the cell cycle regulation of molar tissue uncovered multiple levels of

disregulation at the G1 phase. These included decreased levels of cyclin D1 and cyclin D3 as

well as high levels of the CDK inhibitor p21. These findings suggested that cyclin E1 is

sufficient to push the cell cycle through the G1 phase in trophoblast cells, and highlights a

potential role for p21 in molar development. It will be important in later studies to decipher the

level of interaction and interplay between the D type cyclins, cyclin E1, p21 and p27 in normal

placental development as well as in preeclampsia and molar pregnancy.

145

Figure 6-4: Model of Mtd, cyclin E1 and p27 expression in placentae from severe early onset preeclampsia,

and complete molar pregnancy

Both severe early onset preeclampsia and complete molar pregnancy display increased levels of Mtd expression.

This is associated with increased apoptosis and cell cycle progression. Increased proliferation is likely related to the

elevated levels of the cell cycle activator cyclin E1 (E1), seen in both disorders. In molar pregnancy, the CDK

inhibitor, p27, is decreased. This likely aids in the activating effect of cyclin E1. In contrast, the levels of p27 are

elevated in severe early onset preeclampsia; however, p27 is phosphorylated at the Ser10 site and translocated to the

cytoplasm where it likely does not inhibit cell cycle progression.

146

Together the data presented in this dissertation support our hypothesis that Mtd is an important

regulator of cell death and proliferation in the placenta. Furthermore, this work lends to the

knowledge of how cell cycle regulation may contribute to placentation and suggests that the

balance between Mtd, cyclin E1 and p27 are key to determining the fate of cell cycle progression

in the trophoblast. In conclusion, the work presented in this dissertation has uncovered novel

insights into the regulation of trophoblast apoptosis and cell cycle regulation. Not only is Mtd a

pro-apoptotic regulating molecule, but in the trophoblast it performs a second role in cell cycle

regulation. Importantly, the impact of Mtd on cyclin E1 to promote G1-S transition is a novel

and significant molecular mechanism found to regulate trophoblast cell proliferation in normal

and pathological placentation. Equally important is our identification of molecular differences

that may help to differentiate placental pathologies including early and late onset preeclampsia,

IUGR and molar pregnancy. Cell death and cell cycle regulation are however extremely complex

processes involving a multitude of players. Therefore, although we have uncovered a portion of

the puzzle there are many aspects involved in these events that still need to be addressed further.

147

7 Future Directions – Experimental Design and

Preliminary Data

7.1 Determine if caspase-3 is the connecting link between Mtd and

cyclin E1 through the cleavage of p21 and p27.

The direct mechanism linking Mtd to cyclin E1 expression remains unknown. We hypothesize

that caspase activation, downstream of Mtd, leads to cleavage of p27 and allows for cell cycle

progression. Analysis of the human placental lysates used in our studies revealed the presence of

potential p21 and p27 cleavage products, in addition to the full length protein. Through Western

blot analysis one potential p21 and two potential p27 cleaved products that were identified, were

were differentially expressed across gestation. These bands were consistent with the predicted

size of cleaved p21 and p27, generated by caspases (Figure 7.1a). Moreover, the pattern of the

p27 (c-terminal: 12kd and N-terminal: 15kd, 23kd) bands correlated with the presence of cleaved

caspase-3 over placental development. Examination of cleaved caspase-3 by fluorescence

immunohistochemisty revealed cleaved caspase-3 to be expressed at low levels in the nuclei of

proliferating cytotrophoblast cells in early gestation supporting our hypothesis that caspase

activity may be associated with trophoblast proliferation (Figure 7.1a,b). In future studies it will

be important to verify that theses fragments are truly products of caspase cleavage. This may be

done by assessing the protein cleavage of p21 and p27 from cultured first trimester explants

treated with or without caspase inhibitors. Preliminary studies conducted so far have revealed

that inhibition of caspase-3/7 prevented the cleavage of p27, and this was associated with a

decrease in cyclin D1 expression (Figure 7.1c). The effect of p21/p27 cleavage on cyclin E1

needs to be completed to determine whether caspase activation can potentially link Mtd to cyclin

E1 expression. In addition, it would be informative to assess the rate of proliferation in human

placental explants following treatment with caspase inhibitors. This would help to determine if

caspase activation and p27 cleavage functionally results in the transition from G1-S phase in the

cell cycle.

Future studies comparing the proliferative capabilities of wild type (WT) murine trophoblast

stem cells (TS) and caspase-3 -/- TS cells are also underway to further verify caspase-3

involvement in trophoblast proliferation. This model will ensure effective and specific

knockdown of caspase-3 (Woo et al., 1999). It will be important to first verify caspase-3 specific

148

Figure 7-1 Caspase-3 cleavage of CDK inhibitors in the placenta

a: Protein expression of cleaved caspase-3, p21 (total and cleaved) and p27 (total and cleaved), in total placental

lysates across gestation. Actin was used as an internal control. blue box depicts cleaved caspase-3 and cleaved p27

in early first trimester b: Spatial localization of cleaved caspase-3 (red-top panels) and immunohistochemical dual

staining of cleaved caspase-3 (red) with Ki67 (green) (bottom panels) in early first trimester (5weeks) floating villi.

Nuclei are visualized by DAPI labeled chromatin (blue). Middle and right panels: high magnification of the boxed

area. Circle: cytotrophoblast cell expressing both cleaved caspase-3 and Ki67 (CT: cytotrophoblast; S: stroma; ST:

syncytiotrophoblast). c: Western blot of first trimester placental explants treated with caspase-3/7 specific inhibitor

DEVD, and blotted for total p27 (top panel) and cyclin D1 (bottom panel). Star denotes decreased levels of cleaved

p27 and cyclin D1 in explants treated with DEVD.

149

cleavage of p21 and p27 by comparing their cleavage profiles between caspase3 null and wild

type TS cells. The effect of caspase 3 knockdown on cyclin expression and proliferation could

then be compared between wild type and caspase-3 null TS under proliferating (stem cell

conditions) and differentiating (withdrawal of FGF4) conditions (Tanaka et al., 1998). This could

be done by assessing the expression of cyclin mRNA and protein levels as well as by BrdU

incorporation and FACS cell cycle analysis. To date three TS cell derivations have been

conducted, resulting in 5 wild type cell lines and 3 caspase-3 knock out lines (Figure 7.2).

In light of the data presented in the current dissertation, we hypothesize that targeting p27

cleavage in particular will have a significant impact on cell cycle progression. This could be

investigated further by studying the proliferative capacity of trophoblast cell lines transfected

with p27 constructs with targeted mutations to their caspase-3 cleavage site (ex. D136PSD139S

in p27) (Eymin et al., 1999). This would specifically prevent the cleavage of p27. By comparing

the affect of cleavable and uncleavable p27 on the cell cycle, the significance of p27 cleavage

could be assessed. These studies could be accomplished through BrdU and FACS analysis.

Staining BrdU labeled cells with TO-PRO-3 (T3605, molecular probes) and propidium iodide

(molecular probes) would enable the quantification of both BrdU-labeled and non-labeled cells

so that the percentage of cells actively cycling could be quantified. FACS analysis would also

provide information as to the number of cells in G0, G1,S,G2 and M phase of the cell cycle so

that it could be determined whether cells are arrested or spending less time in a particular phase

of the cell cycle when p27 cleavage is prevented.

7.2 Determine the upstream pathway in preeclampsia leading to the

phosphorylation of p27at Ser10 and its translocation to the

cytoplasm

Determining the mechanism in severe early onset preeclampsia leading to p27 Ser10

phosphorylation and nuclear export will be important to understanding the upstream events

contributing to the disease. Immunoprecipitation studies of p27 with ERK and hKIS in

preeclamptic samples would be useful to determine which kinase interacts with p27 in the

preeclamptic pathology. Additionally, hKIS overexpression studies could be conducted in

trophoblast cell lines to determine if KIS activity leads to p27 phosphorylation and nuclear

export. ERK mediated phosphorylation of p27 could also be verified in trophoblast cell lines by

150

Figure 7-2 Caspase-3 null TS cell derivation

Wild type and caspase-3 deficient TS cells were generated from blastocysts of heterozygous mice carrying a

disrupted Caspase-3 gene. Blastocysts were collected from caspae-3 mutated heterozygous crosses and cultured as

previously described (Tanaka et al., 1998). Cells from each colony were genotyped to identify knockout and Wild

Type (WT) groups.

151

inhibiting MEK activation using the MEK inhibitor PD98059 or U0126 MEK1/2 inhibitor (Cell

Signaling). These studies could be conducted in either BeWo or JEG choriocarcinoma cell lines,

as both cell lines have been shown to have functional ERK pathways (Oufkir et al., 2010). In

addition, estrogens have been shown to cause nuclear export of Ser10 phosphorylated p27 in

early to mid-G1 phase through a mechanism involving the ERK signaling pathway (Foster et al.,

2003). Estrogen levels in preeclampsia should be assessed to determine if estrogen may play a

role upstream in this pathway.

It would also be important to determine if blocking p27 Ser10 phosphorylation in trophoblast

cells would prevent cyclin E1-CDK2 activity and inhibit the positive feedback loop generated by

the excessive cyclin E1 levels in preeclampsia. This could be accomplished by preventing the

kinase activity that targets p27 Ser10 (identified by the experiments outlined above) and

examining cell cycle progression by FACS analysis. Alternatively, this could be accomplished

by knocking down endogenous p27 in trophoblast cell lines and transfecting a p27 construct with

the serine 10 site mutated to an alanine (Ser10Ala). The effect on cyclin E1 mediated CDK2

activity could be assessed by histone phosphorylation assays following cyclin E

immunoprecipitation. In addition, co-immunoprecipitation studies of p27 with cyclin D, CDK4/6

or 14-3-3 may help identify the mode of p27 stabilization.

7.3 Determine the mechanism leading to low p27 expression in molar

tissue and determine if it contributes to increased Mtd and cyclin

E1 expression in the pathology

We hypothesize that the low level of p27 in molar pregnancy may contribute to the

hyperproliferative nature of the disorder. Strategies targeted at elevating p27 in molar cases, or

averting its degradation, may aid in preventing the hyperproliferative component of molar

pathogenesis. Degradation of p27 in the nucleus follows p27 phosphorylation at Thr187 by

CDK2, and involves ubiquitination by the skp2-CKS1 ubiquitin ligase complex (Vlach et al.,

1997;Montagnoli et al., 1999) (Figure 6.5). It would therefore be of interest to examine the

Thr187 phosphorylated and ubiquitinated state of p27 in molar tissue, and to test whether forced

expression of p27 would decrease cyclin E1 and Mtd expression. This could be tested by

blocking p27 ubiquitination using antisense oligos towards skp2 in molar placental explants, and

assessing for CDK2 activation by H1 phosphorylation, as well as cyclin E1 and Mtd expression.

152

REFERENCES

1. (2002). ACOG practice bulletin. Diagnosis and management of preeclampsia and

eclampsia. Number 33, January 2002. American College of Obstetricians and

Gynecologists. Int. J. Gynaecol. Obstet., 77, 67-75.

2. Abrahams,V.M., Kim,Y.M., Straszewski,S.L., Romero,R., and Mor,G. (2004).

Macrophages and apoptotic cell clearance during pregnancy. Am. J. Reprod. Immunol.,

51, 275-282.

3. Adams,P.D. (2001). Regulation of the retinoblastoma tumor suppressor protein by

cyclin/cdks. Biochim. Biophys. Acta, 1471, M123-M133.

4. Alam,A., Cohen,L.Y., Aouad,S., and Sekaly,R.P. (1999). Early activation of caspases

during T lymphocyte stimulation results in selective substrate cleavage in nonapoptotic

cells. J. Exp. Med., 190, 1879-1890.

5. Allaire,A.D., Ballenger,K.A., Wells,S.R., McMahon,M.J., and Lessey,B.A. (2000).

Placental apoptosis in preeclampsia. Obstet. Gynecol., 96, 271-276.

6. Aly,A.S., Khandelwal,M., Zhao,J., Mehmet,A.H., Sammel,M.D., and Parry,S. (2004).

Neutrophils are stimulated by syncytiotrophoblast microvillous membranes to generate

superoxide radicals in women with preeclampsia. Am. J. Obstet. Gynecol., 190, 252-258.

7. Antonsson,B., Montessuit,S., Lauper,S., Eskes,R., and Martinou,J.C. (2000). Bax

oligomerization is required for channel-forming activity in liposomes and to trigger

cytochrome c release from mitochondria. Biochem. J., 345 Pt 2, 271-278.

8. Antonsson,B., Montessuit,S., Sanchez,B., and Martinou,J.C. (2001). Bax is present as a

high molecular weight oligomer/complex in the mitochondrial membrane of apoptotic

cells. J. Biol. Chem., 276, 11615-11623.

9. Arkwright,P.D., Rademacher,T.W., Dwek,R.A., and Redman,C.W. (1993). Pre-eclampsia

is associated with an increase in trophoblast glycogen content and glycogen synthase

activity, similar to that found in hydatidiform moles. J. Clin. Invest, 91, 2744-2753.

10. Arnholdt,H., Meisel,F., Fandrey,K., and Lohrs,U. (1991). Proliferation of villous

trophoblast of the human placenta in normal and abnormal pregnancies. Virchows Arch.

B Cell Pathol. Incl. Mol. Pathol., 60, 365-372.

11. Asada,M., Yamada,T., Ichijo,H., Delia,D., Miyazono,K., Fukumuro,K., and Mizutani,S.

(1999). Apoptosis inhibitory activity of cytoplasmic p21(Cip1/WAF1) in monocytic

differentiation. EMBO J., 18, 1223-1234.

12. Ashton,S.V., Whitley,G.S., Dash,P.R., Wareing,M., Crocker,I.P., Baker,P.N., and

Cartwright,J.E. (2005). Uterine spiral artery remodeling involves endothelial apoptosis

153

induced by extravillous trophoblasts through Fas/FasL interactions. Arterioscler. Thromb.

Vasc. Biol., 25, 102-108.

13. Axt-Fliedner,R., Friedrich,M., Kordina,A., Wasemann,C., Mink,D., Reitnauer,K., and

Schmidt,W. (2001). The immunolocalization of Bcl-2 in human term placenta. Clin. Exp.

Obstet. Gynecol., 28, 144-147.

14. Baczyk,D., Drewlo,S., Proctor,L., Dunk,C., Lye,S., and Kingdom,J. (2009). Glial cell

missing-1 transcription factor is required for the differentiation of the human trophoblast.

Cell Death. Differ., 16, 719-727.

15. Bailly,E., Pines,J., Hunter,T., and Bornens,M. (1992). Cytoplasmic accumulation of

cyclin B1 in human cells: association with a detergent-resistant compartment and with the

centrosome. J. Cell Sci., 101 ( Pt 3), 529-545.

16. Bamberger,A., Sudahl,S., Bamberger,C.M., Schulte,H.M., and Loning,T. (1999).

Expression patterns of the cell-cycle inhibitor p27 and the cell-cycle promoter cyclin E in

the human placenta throughout gestation: implications for the control of proliferation.

Placenta, 20, 401-406.

17. Bamberger,A.M., Aupers,S., Milde-Langosch,K., and Loning,T. (2003). Expression

pattern of the cell cycle promoter cyclin e in benign extravillous trophoblast and

gestational trophoblastic lesions: correlation with expression of Ki-67. Int. J. Gynecol.

Pathol., 22, 156-161.

18. Barker,D.J. (1998). In utero programming of chronic disease. Clin. Sci. (Lond), 95, 115-

128.

19. Barker,D.J. (2003). The developmental origins of adult disease. Eur. J. Epidemiol., 18,

733-736.

20. Barlow,P.W. and Sherman,M.I. (1972). The biochemistry of differentiation of mouse

trophoblast: studies on polyploidy. J. Embryol. Exp. Morphol., 27, 447-465.

21. Bartholomeusz,G., Wu,Y., Ali,S.M., Xia,W., Kwong,K.Y., Hortobagyi,G., and

Hung,M.C. (2006). Nuclear translocation of the pro-apoptotic Bcl-2 family member Bok

induces apoptosis. Mol. Carcinog., 45, 73-83.

22. Besson,A., Dowdy,S.F., and Roberts,J.M. (2008). CDK inhibitors: cell cycle regulators

and beyond. Dev. Cell, 14, 159-169.

23. Besson,A., Gurian-West,M., Chen,X., Kelly-Spratt,K.S., Kemp,C.J., and Roberts,J.M.

(2006). A pathway in quiescent cells that controls p27Kip1 stability, subcellular

localization, and tumor suppression. Genes Dev., 20, 47-64.

24. Besson,A., Gurian-West,M., Schmidt,A., Hall,A., and Roberts,J.M. (2004). p27Kip1

modulates cell migration through the regulation of RhoA activation. Genes Dev., 18, 862-

876.

154

25. Black,S., Kadyrov,M., Kaufmann,P., Ugele,B., Emans,N., and Huppertz,B. (2004).

Syncytial fusion of human trophoblast depends on caspase 8. Cell Death. Differ., 11, 90-

98.

26. Boehm,M., Yoshimoto,T., Crook,M.F., Nallamshetty,S., True,A., Nabel,G.J., and

Nabel,E.G. (2002). A growth factor-dependent nuclear kinase phosphorylates p27(Kip1)

and regulates cell cycle progression. EMBO J., 21, 3390-3401.

27. Bonnefoy-Berard,N., Aouacheria,A., Verschelde,C., Quemeneur,L., Marcais,A., and

Marvel,J. (2004). Control of proliferation by Bcl-2 family members. Biochim. Biophys.

Acta, 1644, 159-168.

28. Boyd,J.D. and Hamilton,W.J. (1970). The Human Placenta. W.Heffer & Sons:

Cambridge, MA. pp 365.

29. Brady,H.J., Gil-Gomez,G., Kirberg,J., and Berns,A.J. (1996). Bax alpha perturbs T cell

development and affects cell cycle entry of T cells. EMBO J., 15, 6991-7001.

30. Brown,L.M., Lacey,H.A., Baker,P.N., and Crocker,I.P. (2005). E-cadherin in the

assessment of aberrant placental cytotrophoblast turnover in pregnancies complicated by

pre-eclampsia. Histochem. Cell Biol., 124, 499-506.

31. Brunelle,J.K. and Chandel,N.S. (2002). Oxygen deprivation induced cell death: an

update. Apoptosis., 7, 475-482.

32. Bryja,V., Pachernik,J., Vondracek,J., Soucek,K., Cajanek,L., Horvath,V., Holubcova,Z.,

Dvorak,P., and Hampl,A. (2008). Lineage specific composition of cyclin D-

CDK4/CDK6-p27 complexes reveals distinct functions of CDK4, CDK6 and individual

D-type cyclins in differentiating cells of embryonic origin. Cell Prolif., 41, 875-893.

33. Burton,G.J. (2009). Oxygen, the Janus gas; its effects on human placental development

and function. J. Anat., 215, 27-35.

34. Burton,G.J. and Jauniaux,E. (2004). Placental oxidative stress: from miscarriage to

preeclampsia. J. Soc. Gynecol. Investig., 11, 342-352.

35. Burton,G.J., Jauniaux,E., and Watson,A.L. (1999). Maternal arterial connections to the

placental intervillous space during the first trimester of human pregnancy: the Boyd

collection revisited. Am. J. Obstet. Gynecol., 181, 718-724.

36. Caniggia,I., Grisaru-Gravnosky,S., Kuliszewsky,M., Post,M., and Lye,S.J. (1999).

Inhibition of TGF-beta 3 restores the invasive capability of extravillous trophoblasts in

preeclamptic pregnancies. J. Clin. Invest, 103, 1641-1650.

37. Caniggia,I., Mostachfi,H., Winter,J., Gassmann,M., Lye,S.J., Kuliszewski,M., and

Post,M. (2000). Hypoxia-inducible factor-1 mediates the biological effects of oxygen on

human trophoblast differentiation through TGFbeta(3). J. Clin. Invest, 105, 577-587.

155

38. Caniggia,I., Taylor,C.V., Ritchie,J.W., Lye,S.J., and Letarte,M. (1997). Endoglin

regulates trophoblast differentiation along the invasive pathway in human placental

villous explants. Endocrinology, 138, 4977-4988.

39. Caniggia,I. and Winter,J.L. (2002). Adriana and Luisa Castellucci Award lecture 2001.

Hypoxia inducible factor-1: oxygen regulation of trophoblast differentiation in normal

and pre-eclamptic pregnancies--a review. Placenta, 23 Suppl A, S47-S57.

40. Cartwright,J.E., Kenny,L.C., Dash,P.R., Crocker,I.P., Aplin,J.D., Baker,P.N., and

Whitley,G.S. (2002). Trophoblast invasion of spiral arteries: a novel in vitro model.

Placenta, 23, 232-235.

41. Cartwright,J.E. and Wareing,M. (2006). An in vitro model of trophoblast invasion of

spiral arteries. Methods Mol. Med., 122, 59-74.

42. Cetin,I. and Alvino,G. (2009). Intrauterine growth restriction: implications for placental

metabolism and transport. A review. Placenta, 30 Suppl A, S77-S82.

43. Cetin,I., Foidart,J.M., Miozzo,M., Raun,T., Jansson,T., Tsatsaris,V., Reik,W., Cross,J.,

Hauguel-de-Mouzon,S., Illsley,N., Kingdom,J., and Huppertz,B. (2004). Fetal growth

restriction: a workshop report. Placenta, 25, 753-757.

44. Chen,C.P., Bajoria,R., and Aplin,J.D. (2002). Decreased vascularization and cell

proliferation in placentas of intrauterine growth-restricted fetuses with abnormal

umbilical artery flow velocity waveforms. Am. J. Obstet. Gynecol., 187, 764-769.

45. Cheng,M., Olivier,P., Diehl,J.A., Fero,M., Roussel,M.F., Roberts,J.M., and Sherr,C.J.

(1999). The p21(Cip1) and p27(Kip1) CDK 'inhibitors' are essential activators of cyclin

D-dependent kinases in murine fibroblasts. EMBO J., 18, 1571-1583.

46. Cheung,A.N., Shen,D.H., Khoo,U.S., Wong,L.C., and Ngan,H.Y. (1998).

p21WAF1/CIP1 expression in gestational trophoblastic disease: correlation with

clinicopathological parameters, and Ki67 and p53 gene expression. J. Clin. Pathol., 51,

159-162.

47. Cheung,A.N., Srivastava,G., Chung,L.P., Ngan,H.Y., Man,T.K., Liu,Y.T., Chen,W.Z.,

Collins,R.J., Wong,L.C., and Ma,H.K. (1994). Expression of the p53 gene in

trophoblastic cells in hydatidiform moles and normal human placentas. J. Reprod. Med.,

39, 223-227.

48. Chiang,M.H., Liang,F.Y., Chen,C.P., Chang,C.W., Cheong,M.L., Wang,L.J., Liang,C.Y.,

Lin,F.Y., Chou,C.C., and Chen,H. (2009). Mechanism of hypoxia-induced GCM1

degradation: implications for the pathogenesis of preeclampsia. J. Biol. Chem., 284,

17411-17419.

49. Child,E.S. and Mann,D.J. (2006). The intricacies of p21 phosphorylation: protein/protein

interactions, subcellular localization and stability. Cell Cycle, 5, 1313-1319.

156

50. Chiu,P.M., Ngan,Y.S., Khoo,U.S., and Cheung,A.N. (2001). Apoptotic activity in

gestational trophoblastic disease correlates with clinical outcome: assessment by the

caspase-related M30 CytoDeath antibody. Histopathology, 38, 243-249.

51. Connor,M.K., Kotchetkov,R., Cariou,S., Resch,A., Lupetti,R., Beniston,R.G.,

Melchior,F., Hengst,L., and Slingerland,J.M. (2003). CRM1/Ran-mediated nuclear

export of p27(Kip1) involves a nuclear export signal and links p27 export and

proteolysis. Mol. Biol. Cell, 14, 201-213.

52. Coverley,D., Laman,H., and Laskey,R.A. (2002). Distinct roles for cyclins E and A

during DNA replication complex assembly and activation. Nat. Cell Biol., 4, 523-528.

53. Crocker,I. (2007). Gabor Than Award Lecture 2006: pre-eclampsia and villous

trophoblast turnover: perspectives and possibilities. Placenta, 28 Suppl A, S4-13.

54. Cross,J.C., Werb,Z., and Fisher,S.J. (1994). Implantation and the placenta: key pieces of

the development puzzle. Science, 266, 1508-1518.

55. Damsky,C.H., Librach,C., Lim,K.H., Fitzgerald,M.L., McMaster,M.T., Janatpour,M.,

Zhou,Y., Logan,S.K., and Fisher,S.J. (1994). Integrin switching regulates normal

trophoblast invasion. Development, 120, 3657-3666.

56. Danihel,L., Gomolcak,P., Korbel,M., Pruzinec,J., Vojtassak,J., Janik,P., and Babal,P.

(2002). Expression of proliferation and apoptotic markers in human placenta during

pregnancy. Acta Histochem., 104, 335-338.

57. De Falco,M., De Luca,L., Acanfora,F., Cavallotti,I., Cottone,G., Laforgia,V., De Luca,B.,

Baldi,A., and De Luca,A. (2001). Alteration of the Bcl-2:Bax ratio in the placenta as

pregnancy proceeds. Histochem. J., 33, 421-425.

58. De Falco,M., Fedele,V., Cobellis,L., Mastrogiacomo,A., Giraldi,D., Leone,S., De

Luca,L., Laforgia,V., and De Luca,A. (2004). Pattern of expression of cyclin D1/CDK4

complex in human placenta during gestation. Cell Tissue Res., 317, 187-194.

59. Degterev,A., Boyce,M., and Yuan,J. (2003). A decade of caspases. Oncogene, 22, 8543-

8567.

60. Dejean,L.M., Martinez-Caballero,S., Guo,L., Hughes,C., Teijido,O., Ducret,T., Ichas,F.,

Korsmeyer,S.J., Antonsson,B., Jonas,E.A., and Kinnally,K.W. (2005). Oligomeric Bax is

a component of the putative cytochrome c release channel MAC, mitochondrial

apoptosis-induced channel. Mol. Biol. Cell, 16, 2424-2432.

61. DeLoia,J.A., Burlingame,J.M., and Krasnow,J.S. (1997). Differential expression of G1

cyclins during human placentogenesis. Placenta, 18, 9-16.

62. Deng,X., Mercer,S.E., Shah,S., Ewton,D.Z., and Friedman,E. (2004). The cyclin-

dependent kinase inhibitor p27Kip1 is stabilized in G(0) by Mirk/dyrk1B kinase. J. Biol.

Chem., 279, 22498-22504.

157

63. DiFederico,E., Genbacev,O., and Fisher,S.J. (1999). Preeclampsia is associated with

widespread apoptosis of placental cytotrophoblasts within the uterine wall. Am. J.

Pathol., 155, 293-301.

64. Du,C., Fang,M., Li,Y., Li,L., and Wang,X. (2000). Smac, a mitochondrial protein that

promotes cytochrome c-dependent caspase activation by eliminating IAP inhibition. Cell,

102, 33-42.

65. Dulic,V., Lees,E., and Reed,S.I. (1992). Association of human cyclin E with a periodic

G1-S phase protein kinase. Science, 257, 1958-1961.

66. Dunk,C., Petkovic,L., Baczyk,D., Rossant,J., Winterhager,E., and Lye,S. (2003). A novel

in vitro model of trophoblast-mediated decidual blood vessel remodeling. Lab Invest, 83,

1821-1828.

67. Dyson,N. (1998). The regulation of E2F by pRB-family proteins. Genes Dev., 12, 2245-

2262.

68. Endl,E. and Gerdes,J. (2000). The Ki-67 protein: fascinating forms and an unknown

function. Exp. Cell Res., 257, 231-237.

69. Everett,R.B. and MacDonald,P.C. (1979). Endocrinology of the placenta. Annu. Rev.

Med., 30, 473-488.

70. Eymin,B., Sordet,O., Droin,N., Munsch,B., Haugg,M., Van de,C.M., Vandenabeele,P.,

and Solary,E. (1999). Caspase-induced proteolysis of the cyclin-dependent kinase

inhibitor p27Kip1 mediates its anti-apoptotic activity. Oncogene, 18, 4839-4847.

71. Feinstein-Rotkopf,Y. and Arama,E. (2009). Can't live without them, can live with them:

roles of caspases during vital cellular processes. Apoptosis., 14, 980-995.

72. Ferrazzi,E., Bozzo,M., Rigano,S., Bellotti,M., Morabito,A., Pardi,G., Battaglia,F.C., and

Galan,H.L. (2002). Temporal sequence of abnormal Doppler changes in the peripheral

and central circulatory systems of the severely growth-restricted fetus. Ultrasound

Obstet. Gynecol., 19, 140-146.

73. Finucane,D.M., Bossy-Wetzel,E., Waterhouse,N.J., Cotter,T.G., and Green,D.R. (1999).

Bax-induced caspase activation and apoptosis via cytochrome c release from

mitochondria is inhibitable by Bcl-xL. J. Biol. Chem., 274, 2225-2233.

74. Fisher,R.A., Nucci,M.R., Thaker,H.M., Weremowicz,S., Genest,D.R., and

Castrillon,D.H. (2004). Complete hydatidiform mole retaining a chromosome 11 of

maternal origin: molecular genetic analysis of a case. Mod. Pathol., 17, 1155-1160.

75. Fisher,S.J., Cui,T.Y., Zhang,L., Hartman,L., Grahl,K., Zhang,G.Y., Tarpey,J., and

Damsky,C.H. (1989). Adhesive and degradative properties of human placental

cytotrophoblast cells in vitro. J. Cell Biol., 109, 891-902.

158

76. Fong,P.Y., Xue,W.C., Ngan,H.Y., Chan,K.Y., Khoo,U.S., Tsao,S.W., Chiu,P.M.,

Man,L.S., and Cheung,A.N. (2005). Mcl-1 expression in gestational trophoblastic disease

correlates with clinical outcome: a differential expression study. Cancer, 103, 268-276.

77. Fong,P.Y., Xue,W.C., Ngan,H.Y., Chiu,P.M., Chan,K.Y., Tsao,S.W., and Cheung,A.N.

(2006). Caspase activity is downregulated in choriocarcinoma: a cDNA array differential

expression study. J. Clin. Pathol., 59, 179-183.

78. Foster,J.S., Fernando,R.I., Ishida,N., Nakayama,K.I., and Wimalasena,J. (2003).

Estrogens down-regulate p27Kip1 in breast cancer cells through Skp2 and through

nuclear export mediated by the ERK pathway. J. Biol. Chem., 278, 41355-41366.

79. Fuentes-Prior,P. and Salvesen,G.S. (2004). The protein structures that shape caspase

activity, specificity, activation and inhibition. Biochem. J., 384, 201-232.

80. Fujise,K., Zhang,D., Liu,J., and Yeh,E.T. (2000). Regulation of apoptosis and cell cycle

progression by MCL1. Differential role of proliferating cell nuclear antigen. J. Biol.

Chem., 275, 39458-39465.

81. Fujita,N., Sato,S., Katayama,K., and Tsuruo,T. (2002). Akt-dependent phosphorylation of

p27Kip1 promotes binding to 14-3-3 and cytoplasmic localization. J. Biol. Chem., 277,

28706-28713.

82. Fukunaga,M. (2004). Immunohistochemical characterization of cyclin E and p27KIP1

expression in early hydatidiform moles. Int. J. Gynecol. Pathol., 23, 259-264.

83. Galan,A., O'Connor,J.E., Valbuena,D., Herrer,R., Remohi,J., Pampfer,S., Pellicer,A., and

Simon,C. (2000). The human blastocyst regulates endometrial epithelial apoptosis in

embryonic adhesion. Biol. Reprod., 63, 430-439.

84. Gao,S., Fu,W., Durrenberger,M., De Geyter,C., and Zhang,H. (2005). Membrane

translocation and oligomerization of hBok are triggered in response to apoptotic stimuli

and Bnip3. Cell Mol. Life Sci., 62, 1015-1024.

85. Genbacev,O., DiFederico,E., McMaster,M., and Fisher,S.J. (1999). Invasive

cytotrophoblast apoptosis in pre-eclampsia. Hum. Reprod., 14 Suppl 2, 59-66.

86. Genbacev,O., Joslin,R., Damsky,C.H., Polliotti,B.M., and Fisher,S.J. (1996). Hypoxia

alters early gestation human cytotrophoblast differentiation/invasion in vitro and models

the placental defects that occur in preeclampsia. J. Clin. Invest, 97, 540-550.

87. Genbacev,O., McMaster,M.T., and Fisher,S.J. (2000). A repertoire of cell cycle

regulators whose expression is coordinated with human cytotrophoblast differentiation.

Am. J. Pathol., 157, 1337-1351.

88. Genbacev,O., Zhou,Y., Ludlow,J.W., and Fisher,S.J. (1997). Regulation of human

placental development by oxygen tension. Science, 277, 1669-1672.

159

89. Geng,Y., Eaton,E.N., Picon,M., Roberts,J.M., Lundberg,A.S., Gifford,A., Sardet,C., and

Weinberg,R.A. (1996). Regulation of cyclin E transcription by E2Fs and retinoblastoma

protein. Oncogene, 12, 1173-1180.

90. Geng,Y., Yu,Q., Sicinska,E., Das,M., Schneider,J.E., Bhattacharya,S., Rideout,W.M.,

Bronson,R.T., Gardner,H., and Sicinski,P. (2003). Cyclin E ablation in the mouse. Cell,

114, 431-443.

91. Geng,Y., Yu,Q., Whoriskey,W., Dick,F., Tsai,K.Y., Ford,H.L., Biswas,D.K.,

Pardee,A.B., Amati,B., Jacks,T., Richardson,A., Dyson,N., and Sicinski,P. (2001).

Expression of cyclins E1 and E2 during mouse development and in neoplasia. Proc. Natl.

Acad. Sci. U. S. A, 98, 13138-13143.

92. Gerretsen,G., Huisjes,H.J., and Elema,J.D. (1981). Morphological changes of the spiral

arteries in the placental bed in relation to pre-eclampsia and fetal growth retardation. Br.

J. Obstet. Gynaecol., 88, 876-881.

93. Graham,C.H. and Lala,P.K. (1991). Mechanism of control of trophoblast invasion in situ.

J. Cell Physiol, 148, 228-234.

94. Graham,C.H. and Lala,P.K. (1992). Mechanisms of placental invasion of the uterus and

their control. Biochem. Cell Biol., 70, 867-874.

95. Green,D.R. and Reed,J.C. (1998). Mitochondria and apoptosis. Science, 281, 1309-1312.

96. Gross,A., McDonnell,J.M., and Korsmeyer,S.J. (1999). BCL-2 family members and the

mitochondria in apoptosis. Genes Dev., 13, 1899-1911.

97. Gurel,D., Ozer,E., Altunyurt,S., Guclu,S., and Demir,N. (2003). Expression of IGR-IR

and VEGF and trophoblastic proliferative activity in placentas from pregnancies

complicated by IUGR. Pathol. Res. Pract., 199, 803-809.

98. Habli,M., Levine,R.J., Qian,C., and Sibai,B. (2007). Neonatal outcomes in pregnancies

with preeclampsia or gestational hypertension and in normotensive pregnancies that

delivered at 35, 36, or 37 weeks of gestation. Am. J. Obstet. Gynecol., 197, 406-407.

99. Halperin,R., Peller,S., Sandbank,J., Bukovsky,I., and Schneider,D. (2000). Expression of

the p53 gene and apoptosis in gestational trophoblastic disease. Placenta, 21, 58-62.

100. Hansen,A.R., Collins,M.H., Genest,D., Heller,D., Shen-Schwarz,S., Banagon,P.,

Allred,E.N., and Leviton,A. (2000). Very low birthweight placenta: clustering of

morphologic characteristics. Pediatr. Dev. Pathol., 3, 431-438.

101. Hauth,J.C., Ewell,M.G., Levine,R.J., Esterlitz,J.R., Sibai,B., Curet,L.B., Catalano,P.M.,

and Morris,C.D. (2000). Pregnancy outcomes in healthy nulliparas who developed

hypertension. Calcium for Preeclampsia Prevention Study Group. Obstet. Gynecol., 95,

24-28.

160

102. Heazell,A.E. and Crocker,I.P. (2008a). Live and let die - regulation of villous trophoblast

apoptosis in normal and abnormal pregnancies. Placenta, 29, 772-783.

103. Heazell,A.E., Lacey,H.A., Jones,C.J., Huppertz,B., Baker,P.N., and Crocker,I.P. (2008b).

Effects of oxygen on cell turnover and expression of regulators of apoptosis in human

placental trophoblast. Placenta, 29, 175-186.

104. Hsu,S.Y. and Hsueh,A.J. (2000). Tissue-specific Bcl-2 protein partners in apoptosis: An

ovarian paradigm. Physiol Rev., 80, 593-614.

105. Hsu,S.Y., Kaipia,A., McGee,E., Lomeli,M., and Hsueh,A.J. (1997). Bok is a pro-

apoptotic Bcl-2 protein with restricted expression in reproductive tissues and

heterodimerizes with selective anti-apoptotic Bcl-2 family members. Proc. Natl. Acad.

Sci. U. S. A, 94, 12401-12406.

106. Hu,R., Zhou,S., and Li,X. (2006). Altered Bcl-2 and Bax expression is associated with

cultured first trimester human cytotrophoblasts apoptosis induced by hypoxia. Life Sci.,

79, 351-355.

107. Hubel,C.A. (1999). Oxidative stress in the pathogenesis of preeclampsia. Proc. Soc. Exp.

Biol. Med., 222, 222-235.

108. Hung,T.H., Skepper,J.N., and Burton,G.J. (2001). In vitro ischemia-reperfusion injury in

term human placenta as a model for oxidative stress in pathological pregnancies. Am. J.

Pathol., 159, 1031-1043.

109. Hung,T.H., Skepper,J.N., Charnock-Jones,D.S., and Burton,G.J. (2002). Hypoxia-

reoxygenation: a potent inducer of apoptotic changes in the human placenta and possible

etiological factor in preeclampsia. Circ. Res., 90, 1274-1281.

110. Huppertz,B., Bartz,C., and Kokozidou,M. (2006). Trophoblast fusion: Fusogenic

proteins, syncytins and ADAMs, and other prerequisites for syncytial fusion. Micron..

111. Huppertz,B., Frank,H.G., Kingdom,J.C., Reister,F., and Kaufmann,P. (1998). Villous

cytotrophoblast regulation of the syncytial apoptotic cascade in the human placenta.

Histochem. Cell Biol., 110, 495-508.

112. Huppertz,B., Frank,H.G., Reister,F., Kingdom,J., Korr,H., and Kaufmann,P. (1999).

Apoptosis cascade progresses during turnover of human trophoblast: analysis of villous

cytotrophoblast and syncytial fragments in vitro. Lab Invest, 79, 1687-1702.

113. Huppertz,B., Kingdom,J., Caniggia,I., Desoye,G., Black,S., Korr,H., and Kaufmann,P.

(2003). Hypoxia favours necrotic versus apoptotic shedding of placental

syncytiotrophoblast into the maternal circulation. Placenta, 24, 181-190.

114. Huppertz,B. and Kingdom,J.C. (2004). Apoptosis in the trophoblast--role of apoptosis in

placental morphogenesis. J. Soc. Gynecol. Investig., 11, 353-362.

161

115. Ietta,F., Wu,Y., Romagnoli,R., Soleymanlou,N., Orsini,B., Zamudio,S., Paulesu,L., and

Caniggia,I. (2007). Oxygen regulation of macrophage migration inhibitory factor in

human placenta. Am. J. Physiol Endocrinol. Metab, 292, E272-E280.

116. Ietta,F., Wu,Y., Winter,J., Xu,J., Wang,J., Post,M., and Caniggia,I. (2006). Dynamic

HIF1A regulation during human placental development. Biol. Reprod., 75, 112-121.

117. Inohara,N., Ekhterae,D., Garcia,I., Carrio,R., Merino,J., Merry,A., Chen,S., and

Nunez,G. (1998). Mtd, a novel Bcl-2 family member activates apoptosis in the absence of

heterodimerization with Bcl-2 and Bcl-XL. J. Biol. Chem., 273, 8705-8710.

118. Irving,J.A. and Lala,P.K. (1995). Functional role of cell surface integrins on human

trophoblast cell migration: regulation by TGF-beta, IGF-II, and IGFBP-1. Exp. Cell Res.,

217, 419-427.

119. Ishida,N., Kitagawa,M., Hatakeyama,S., and Nakayama,K. (2000). Phosphorylation at

serine 10, a major phosphorylation site of p27(Kip1), increases its protein stability. J.

Biol. Chem., 275, 25146-25154.

120. Ishihara,N., Matsuo,H., Murakoshi,H., Laoag-Fernandez,J.B., Samoto,T., and Maruo,T.

(2002). Increased apoptosis in the syncytiotrophoblast in human term placentas

complicated by either preeclampsia or intrauterine growth retardation. Am. J. Obstet.

Gynecol., 186, 158-166.

121. Jackman,M., Kubota,Y., den Elzen,N., Hagting,A., and Pines,J. (2002). Cyclin A- and

cyclin E-Cdk complexes shuttle between the nucleus and the cytoplasm. Mol. Biol. Cell,

13, 1030-1045.

122. Jaffe,R., Jauniaux,E., and Hustin,J. (1997). Maternal circulation in the first-trimester

human placenta--myth or reality? Am. J. Obstet. Gynecol., 176, 695-705.

123. Jamil,S., Sobouti,R., Hojabrpour,P., Raj,M., Kast,J., and Duronio,V. (2005). A

proteolytic fragment of Mcl-1 exhibits nuclear localization and regulates cell growth by

interaction with Cdk1. Biochem. J., 387, 659-667.

124. Jauniaux,E., Watson,A.L., Hempstock,J., Bao,Y.P., Skepper,J.N., and Burton,G.J.

(2000). Onset of maternal arterial blood flow and placental oxidative stress. A possible

factor in human early pregnancy failure. Am. J. Pathol., 157, 2111-2122.

125. Jeschke,U., Schiessl,B., Mylonas,I., Kunze,S., Kuhn,C., Schulze,S., Friese,K., and

Mayr,D. (2006). Expression of the proliferation marker Ki-67 and of p53 tumor protein in

trophoblastic tissue of preeclamptic, HELLP, and intrauterine growth-restricted

pregnancies. Int. J. Gynecol. Pathol., 25, 354-360.

126. Johansen,M., Redman,C.W., Wilkins,T., and Sargent,I.L. (1999). Trophoblast

deportation in human pregnancy--its relevance for pre-eclampsia. Placenta, 20, 531-539.

162

127. Kajantie,E., Eriksson,J.G., Osmond,C., Thornburg,K., and Barker,D.J. (2009). Pre-

eclampsia is associated with increased risk of stroke in the adult offspring: the Helsinki

birth cohort study. Stroke, 40, 1176-1180.

128. Kale,A., Soylemez,F., and Ensari,A. (2001). Expressions of proliferation markers (Ki-67,

proliferating cell nuclear antigen, and silver-staining nucleolar organizer regions) and of

p53 tumor protein in gestational trophoblastic disease. Am. J. Obstet. Gynecol., 184, 567-

574.

129. Kaufmann,P. (1982). Development and differentiation of the human placental villous

tree. Bibl. Anat., 29-39.

130. Kim,K.R., Park,B.H., Hong,Y.O., Kwon,H.C., and Robboy,S.J. (2009). The villous

stromal constituents of complete hydatidiform mole differ histologically in very early

pregnancy from the normally developing placenta. Am. J. Surg. Pathol., 33, 176-185.

131. Kim,M.J., Kim,K.R., Ro,J.Y., Lage,J.M., and Lee,H.I. (2006). Diagnostic and

pathogenetic significance of increased stromal apoptosis and incomplete vasculogenesis

in complete hydatidiform moles in very early pregnancy periods. Am. J. Surg. Pathol.,

30, 362-369.

132. Kim,Y.T., Cho,N.H., Ko,J.H., Yang,W.I., Kim,J.W., Choi,E.K., and Lee,S.H. (2000).

Expression of cyclin E in placentas with hydropic change and gestational trophoblastic

diseases: implications for the malignant transformation of trophoblasts. Cancer, 89, 673-

679.

133. Kingdom,J.C. and Kaufmann,P. (1997). Oxygen and placental villous development:

origins of fetal hypoxia. Placenta, 18, 613-621.

134. Kipp,B.R., Ketterling,R.P., Oberg,T.N., Cousin,M.A., Plagge,A.M., Wiktor,A.E.,

Ihrke,J.M., Meyers,C.H., Morice,W.G., Halling,K.C., and Clayton,A.C. (2010).

Comparison of fluorescence in situ hybridization, p57 immunostaining, flow cytometry,

and digital image analysis for diagnosing molar and nonmolar products of conception.

Am. J. Clin. Pathol., 133, 196-204.

135. Knoblich,J.A., Sauer,K., Jones,L., Richardson,H., Saint,R., and Lehner,C.F. (1994).

Cyclin E controls S phase progression and its down-regulation during Drosophila

embryogenesis is required for the arrest of cell proliferation. Cell, 77, 107-120.

136. Knudson,C.M., Johnson,G.M., Lin,Y., and Korsmeyer,S.J. (2001). Bax accelerates

tumorigenesis in p53-deficient mice. Cancer Res., 61, 659-665.

137. Koff,A., Giordano,A., Desai,D., Yamashita,K., Harper,J.W., Elledge,S., Nishimoto,T.,

Morgan,D.O., Franza,B.R., and Roberts,J.M. (1992). Formation and activation of a cyclin

E-cdk2 complex during the G1 phase of the human cell cycle. Science, 257, 1689-1694.

138. Koga,K., Osuga,Y., Tajima,T., Hirota,Y., Igarashi,T., Fujii,T., Yano,T., and Taketani,Y.

(2009). Elevated serum soluble fms-like tyrosine kinase 1 (sFlt1) level in women with

hydatidiform mole. Fertil. Steril..

163

139. Korgun,E.T., Celik-Ozenci,C., Acar,N., Cayli,S., Desoye,G., and Demir,R. (2006).

Location of cell cycle regulators cyclin B1, cyclin A, PCNA, Ki67 and cell cycle

inhibitors p21, p27 and p57 in human first trimester placenta and deciduas. Histochem.

Cell Biol..

140. Kozar,K., Ciemerych,M.A., Rebel,V.I., Shigematsu,H., Zagozdzon,A., Sicinska,E.,

Geng,Y., Yu,Q., Bhattacharya,S., Bronson,R.T., Akashi,K., and Sicinski,P. (2004).

Mouse development and cell proliferation in the absence of D-cyclins. Cell, 118, 477-

491.

141. Kurman,R.J. (1991a). Pathology of trophoblast. Monogr Pathol., 195-227.

142. Kurman,R.J. (1991b). The morphology, biology, and pathology of intermediate

trophoblast: a look back to the present. Hum. Pathol., 22, 847-855.

143. Labaer,J., Garrett,M.D., Stevenson,L.F., Slingerland,J.M., Sandhu,C., Chou,H.S.,

Fattaey,A., and Harlow,E. (1997). New functional activities for the p21 family of CDK

inhibitors. Genes Dev., 11, 847-862.

144. Lamkanfi,M., Festjens,N., Declercq,W., Berghe,T.V., and Vandenabeele,P. (2007).

Caspases in cell survival, proliferation and differentiation. Cell Death. Differ., 14, 44-55.

145. Launay,S., Hermine,O., Fontenay,M., Kroemer,G., Solary,E., and Garrido,C. (2005).

Vital functions for lethal caspases. Oncogene, 24, 5137-5148.

146. Lea,R.G., Riley,S.C., Antipatis,C., Hannah,L., Ashworth,C.J., Clark,D.A., and

Critchley,H.O. (1999). Cytokines and the regulation of apoptosis in reproductive tissues:

a review. Am. J. Reprod. Immunol., 42, 100-109.

147. Leung,D.N., Smith,S.C., To,K.F., Sahota,D.S., and Baker,P.N. (2001). Increased

placental apoptosis in pregnancies complicated by preeclampsia. Am. J. Obstet. Gynecol.,

184, 1249-1250.

148. Levine,R.J., Maynard,S.E., Qian,C., Lim,K.H., England,L.J., Yu,K.F., Schisterman,E.F.,

Thadhani,R., Sachs,B.P., Epstein,F.H., Sibai,B.M., Sukhatme,V.P., and Karumanchi,S.A.

(2004). Circulating angiogenic factors and the risk of preeclampsia. N. Engl. J. Med.,

350, 672-683.

149. Levy,R. and Nelson,D.M. (2000). To be, or not to be, that is the question. Apoptosis in

human trophoblast. Placenta, 21, 1-13.

150. Levy,R., Smith,S.D., Yusuf,K., Huettner,P.C., Kraus,F.T., Sadovsky,Y., and

Nelson,D.M. (2002). Trophoblast apoptosis from pregnancies complicated by fetal

growth restriction is associated with enhanced p53 expression. Am. J. Obstet. Gynecol.,

186, 1056-1061.

151. Lew,D.J., Dulic,V., and Reed,S.I. (1991). Isolation of three novel human cyclins by

rescue of G1 cyclin (Cln) function in yeast. Cell, 66, 1197-1206.

164

152. Li,H.W., Tsao,S.W., and Cheung,A.N. (2002). Current understandings of the molecular

genetics of gestational trophoblastic diseases. Placenta, 23, 20-31.

153. Liang,J., Shao,S.H., Xu,Z.X., Hennessy,B., Ding,Z., Larrea,M., Kondo,S., Dumont,D.J.,

Gutterman,J.U., Walker,C.L., Slingerland,J.M., and Mills,G.B. (2007). The energy

sensing LKB1-AMPK pathway regulates p27(kip1) phosphorylation mediating the

decision to enter autophagy or apoptosis. Nat. Cell Biol., 9, 218-224.

154. Liu,Y., Yeh,N., Zhu,X.H., Leversha,M., Cordon-Cardo,C., Ghossein,R., Singh,B.,

Holland,E., and Koff,A. (2007). Somatic cell type specific gene transfer reveals a tumor-

promoting function for p21(Waf1/Cip1). EMBO J., 26, 4683-4693.

155. Livak,K.J. and Schmittgen,T.D. (2001). Analysis of relative gene expression data using

real-time quantitative PCR and the 2(-Delta Delta C(T)) Method. Methods, 25, 402-408.

156. Lundberg,A.S. and Weinberg,R.A. (1998). Functional inactivation of the retinoblastoma

protein requires sequential modification by at least two distinct cyclin-cdk complexes.

Mol. Cell Biol., 18, 753-761.

157. MacAuley,A., Cross,J.C., and Werb,Z. (1998). Reprogramming the cell cycle for

endoreduplication in rodent trophoblast cells. Mol. Biol. Cell, 9, 795-807.

158. MacPhee,D.J., Mostachfi,H., Han,R., Lye,S.J., Post,M., and Caniggia,I. (2001). Focal

adhesion kinase is a key mediator of human trophoblast development. Lab Invest, 81,

1469-1483.

159. Maddika,S., Ande,S.R., Panigrahi,S., Paranjothy,T., Weglarczyk,K., Zuse,A.,

Eshraghi,M., Manda,K.D., Wiechec,E., and Los,M. (2007). Cell survival, cell death and

cell cycle pathways are interconnected: implications for cancer therapy. Drug Resist.

Updat., 10, 13-29.

160. Massague,J. (1998). TGF-beta signal transduction. Annu. Rev. Biochem., 67, 753-791.

161. Massardier,J., Golfier,F., Journet,D., Frappart,L., Zalaquett,M., Schott,A.M.,

Lenoir,V.T., Dupuis,O., Hajri,T., and Raudrant,D. (2009). Twin pregnancy with complete

hydatidiform mole and coexistent fetus: obstetrical and oncological outcomes in a series

of 14 cases. Eur. J. Obstet. Gynecol. Reprod. Biol., 143, 84-87.

162. Matsui,H., Iitsuka,Y., Yamazawa,K., Tanaka,N., Mitsuhashi,A., Seki,K., and Sekiya,S.

(2003). Criteria for initiating chemotherapy in patients after evacuation of hydatidiform

mole. Tumour. Biol., 24, 140-146.

163. Maulik,D. (2006). Management of fetal growth restriction: an evidence-based approach.

Clin. Obstet. Gynecol., 49, 320-334.

164. Mazur,M.T. and Kurman,R.J. (1994). Gestational trophoblastic disease and related

lesions. Blaustein's pathology of the female genital tract, . New York: Springer-Verlag,

pp. 1049-1093.

165

165. McAllister,S.S., Becker-Hapak,M., Pintucci,G., Pagano,M., and Dowdy,S.F. (2003).

Novel p27(kip1) C-terminal scatter domain mediates Rac-dependent cell migration

independent of cell cycle arrest functions. Mol. Cell Biol., 23, 216-228.

166. McConnell,T.G., Murphy,K.M., Hafez,M., Vang,R., and Ronnett,B.M. (2009a).

Diagnosis and subclassification of hydatidiform moles using p57 immunohistochemistry

and molecular genotyping: validation and prospective analysis in routine and consultation

practice settings with development of an algorithmic approach. Am. J. Surg. Pathol., 33,

805-817.

167. McConnell,T.G., Norris-Kirby,A., Hagenkord,J.M., Ronnett,B.M., and Murphy,K.M.

(2009b). Complete hydatidiform mole with retained maternal chromosomes 6 and 11.

Am. J. Surg. Pathol., 33, 1409-1415.

168. McKenzie,P.P., Foster,J.S., House,S., Bukovsky,A., Caudle,M.R., and Wimalasena,J.

(1998). Expression of G1 cyclins and cyclin-dependent kinase-2 activity during terminal

differentiation of cultured human trophoblast. Biol. Reprod., 58, 1283-1289.

169. Miyashita,T., Krajewski,S., Krajewska,M., Wang,H.G., Lin,H.K., Liebermann,D.A.,

Hoffman,B., and Reed,J.C. (1994). Tumor suppressor p53 is a regulator of bcl-2 and bax

gene expression in vitro and in vivo. Oncogene, 9, 1799-1805.

170. Montagnoli,A., Fiore,F., Eytan,E., Carrano,A.C., Draetta,G.F., Hershko,A., and

Pagano,M. (1999). Ubiquitination of p27 is regulated by Cdk-dependent phosphorylation

and trimeric complex formation. Genes Dev., 13, 1181-1189.

171. Moore and Persaud (1998). The Developing Human: Clinically Oriented Embryology.

172. Morgan,D.O. (1997). Cyclin-dependent kinases: engines, clocks, and microprocessors.

Annu. Rev. Cell Dev. Biol., 13, 261-291.

173. Moroy,T. and Geisen,C. (2004). Cyclin E. Int. J. Biochem. Cell Biol., 36, 1424-1439.

174. Morrish,D.W., Dakour,J., and Li,H. (1998). Functional regulation of human trophoblast

differentiation. J. Reprod. Immunol., 39, 179-195.

175. Myatt,L. (2006). Placental adaptive responses and fetal programming. J. Physiol, 572,

25-30.

176. Nakamura,S., Okinaka,K., Hirano,I., Ono,T., Sugimoto,Y., Shigeno,K., Fujisawa,S.,

Shinjo,K., and Ohnishi,K. (2008). KIS induces proliferation and the cell cycle

progression through the phosphorylation of p27Kip1 in leukemia cells. Leuk. Res., 32,

1358-1365.

177. Nevins,J.R. (1998). Toward an understanding of the functional complexity of the E2F

and retinoblastoma families. Cell Growth Differ., 9, 585-593.

178. Nevo,O., Many,A., Xu,J., Kingdom,J., Piccoli,E., Zamudio,S., Post,M., Bocking,A.,

Todros,T., and Caniggia,I. (2008). Placental expression of soluble fms-like tyrosine

166

kinase 1 is increased in singletons and twin pregnancies with intrauterine growth

restriction. J. Clin. Endocrinol. Metab, 93, 285-292.

179. Nevo,O., Soleymanlou,N., Wu,Y., Xu,J., Kingdom,J., Many,A., Zamudio,S., and

Caniggia,I. (2006). Increased expression of sFlt-1 in in vivo and in vitro models of

human placental hypoxia is mediated by HIF-1. Am. J. Physiol Regul. Integr. Comp

Physiol, 291, R1085-R1093.

180. Olvera,M., Harris,S., Amezcua,C.A., McCourty,A., Rezk,S., Koo,C., Felix,J.C., and

Brynes,R.K. (2001). Immunohistochemical expression of cell cycle proteins E2F-1, Cdk-

2, Cyclin E, p27(kip1), and Ki-67 in normal placenta and gestational trophoblastic

disease. Mod. Pathol., 14, 1036-1042.

181. Orcy,R.B., Schroeder,S., Martins-Costa,S.H., Ramos,J.G., Schechinger,W., Klein,H.,

Brum,I.S., von Eye,C.H., and Capp,E. (2008). Signalization of Akt/PKB in the placenta,

skeletal muscle and adipose tissue of preeclampsia patients. Gynecol. Obstet. Invest, 66,

231-236.

182. Oufkir,T., Arseneault,M., Sanderson,J.T., and Vaillancourt,C. (2010). The 5-HT(2A)

serotonin receptor enhances cell viability, affects cell cycle progression and activates

MEK-ERK1/2 and JAK2-STAT3 signalling pathways in human choriocarcinoma cell

lines. Placenta.

183. Palmer,J.R. (1994). Advances in the epidemiology of gestational trophoblastic disease. J.

Reprod. Med., 39, 155-162.

184. Panichkul,P.C., Al Hussaini,T.K., Sierra,R., Kashork,C.D., Popek,E.J., Stockton,D.W.,

and Van,d.V., I (2005). Recurrent biparental hydatidiform mole: additional evidence for a

1.1-Mb locus in 19q13.4 and candidate gene analysis. J. Soc. Gynecol. Investig., 12, 376-

383.

185. Paradinas,F.J., Browne,P., Fisher,R.A., Foskett,M., Bagshawe,K.D., and Newlands,E.

(1996). A clinical, histopathological and flow cytometric study of 149 complete moles,

146 partial moles and 107 non-molar hydropic abortions. Histopathology, 28, 101-110.

186. Parisi,T., Beck,A.R., Rougier,N., McNeil,T., Lucian,L., Werb,Z., and Amati,B. (2003).

Cyclins E1 and E2 are required for endoreplication in placental trophoblast giant cells.

EMBO J., 22, 4794-4803.

187. Pollack,R.N. and Divon,M.Y. (1992). Intrauterine growth retardation: definition,

classification, and etiology. Clin. Obstet. Gynecol., 35, 99-107.

188. Polyak,K., Kato,J.Y., Solomon,M.J., Sherr,C.J., Massague,J., Roberts,J.M., and Koff,A.

(1994). p27Kip1, a cyclin-Cdk inhibitor, links transforming growth factor-beta and

contact inhibition to cell cycle arrest. Genes Dev., 8, 9-22.

189. Potgens,A.J., Schmitz,U., Bose,P., Versmold,A., Kaufmann,P., and Frank,H.G. (2002).

Mechanisms of syncytial fusion: a review. Placenta, 23 Suppl A, S107-S113.

167

190. Qiao,S., Nagasaka,T., Harada,T., and Nakashima,N. (1998). p53, Bax and Bcl-2

expression, and apoptosis in gestational trophoblast of complete hydatidiform mole.

Placenta, 19, 361-369.

191. Rajakumar,A. and Conrad,K.P. (2000). Expression, ontogeny, and regulation of hypoxia-

inducible transcription factors in the human placenta. Biol. Reprod., 63, 559-569.

192. Ratts,V.S., Tao,X.J., Webster,C.B., Swanson,P.E., Smith,S.D., Brownbill,P.,

Krajewski,S., Reed,J.C., Tilly,J.L., and Nelson,D.M. (2000). Expression of BCL-2, BAX

and BAK in the trophoblast layer of the term human placenta: a unique model of

apoptosis within a syncytium. Placenta, 21, 361-366.

193. Ray,J., Jurisicova,A., and Caniggia,I. (2008). IFPA Trophoblast Research Award

Lecture: The Dynamic Role of Bcl-2 Family Members in Trophoblast Cell Fate.

Placenta.

194. Ray,J.E., Garcia,J., Jurisicova,A., and Caniggia,I. (2009). Mtd/Bok takes a swing:

proapoptotic Mtd/Bok regulates trophoblast cell proliferation during human placental

development and in preeclampsia. Cell Death. Differ..

195. Redline,R.W. (2008). Placental pathology: a systematic approach with clinical

correlations. Placenta, 29 Suppl A, S86-S91.

196. Redline,R.W. and Patterson,P. (1995). Pre-eclampsia is associated with an excess of

proliferative immature intermediate trophoblast. Hum. Pathol., 26, 594-600.

197. Redman,C.W., Sacks,G.P., and Sargent,I.L. (1999). Preeclampsia: an excessive maternal

inflammatory response to pregnancy. Am. J. Obstet. Gynecol., 180, 499-506.

198. Redman,C.W. and Sargent,I.L. (2000). Placental debris, oxidative stress and pre-

eclampsia. Placenta, 21, 597-602.

199. Redman,C.W. and Sargent,I.L. (2003). Pre-eclampsia, the placenta and the maternal

systemic inflammatory response--a review. Placenta, 24 Suppl A, S21-S27.

200. Roberts,J.M. and Cooper,D.W. (2001). Pathogenesis and genetics of pre-eclampsia.

Lancet, 357, 53-56.

201. Rodesch,F., Simon,P., Donner,C., and Jauniaux,E. (1992). Oxygen measurements in

endometrial and trophoblastic tissues during early pregnancy. Obstet. Gynecol., 80, 283-

285.

202. Rodier,G., Montagnoli,A., Di Marcotullio,L., Coulombe,P., Draetta,G.F., Pagano,M., and

Meloche,S. (2001). p27 cytoplasmic localization is regulated by phosphorylation on

Ser10 and is not a prerequisite for its proteolysis. EMBO J., 20, 6672-6682.

203. Rodriguez,J.M., Glozak,M.A., Ma,Y., and Cress,W.D. (2006). Bok, Bcl-2-related

Ovarian Killer, Is Cell Cycle-regulated and Sensitizes to Stress-induced Apoptosis. J.

Biol. Chem., 281, 22729-22735.

168

204. Romero,R., Nien,J.K., Espinoza,J., Todem,D., Fu,W., Chung,H., Kusanovic,J.P.,

Gotsch,F., Erez,O., Mazaki-Tovi,S., Gomez,R., Edwin,S., Chaiworapongsa,T.,

Levine,R.J., and Karumanchi,S.A. (2008). A longitudinal study of angiogenic (placental

growth factor) and anti-angiogenic (soluble endoglin and soluble vascular endothelial

growth factor receptor-1) factors in normal pregnancy and patients destined to develop

preeclampsia and deliver a small for gestational age neonate. J. Matern. Fetal Neonatal

Med., 21, 9-23.

205. Russo,A.A., Jeffrey,P.D., Patten,A.K., Massague,J., and Pavletich,N.P. (1996). Crystal

structure of the p27Kip1 cyclin-dependent-kinase inhibitor bound to the cyclin A-Cdk2

complex. Nature, 382, 325-331.

206. Sargent,I.L., Germain,S.J., Sacks,G.P., Kumar,S., and Redman,C.W. (2003). Trophoblast

deportation and the maternal inflammatory response in pre-eclampsia. J. Reprod.

Immunol., 59, 153-160.

207. Sebire,N.J. (2010). Histopathological diagnosis of hydatidiform mole: contemporary

features and clinical implications. Fetal Pediatr. Pathol., 29, 1-16.

208. Seckl,M.J., Fisher,R.A., Salerno,G., Rees,H., Paradinas,F.J., Foskett,M., and

Newlands,E.S. (2000). Choriocarcinoma and partial hydatidiform moles. Lancet, 356, 36-

39.

209. Sekimoto,T., Fukumoto,M., and Yoneda,Y. (2004). 14-3-3 suppresses the nuclear

localization of threonine 157-phosphorylated p27(Kip1). EMBO J., 23, 1934-1942.

210. Sgambato,A., Cittadini,A., Faraglia,B., and Weinstein,I.B. (2000). Multiple functions of

p27(Kip1) and its alterations in tumor cells: a review. J. Cell Physiol, 183, 18-27.

211. Sherr,C.J. (1994). G1 phase progression: cycling on cue. Cell, 79, 551-555.

212. Sherr,C.J. (1995). D-type cyclins. Trends Biochem. Sci., 20, 187-190.

213. Sherr,C.J. (1996). Cancer cell cycles. Science, 274, 1672-1677.

214. Sherr,C.J. and Roberts,J.M. (1995). Inhibitors of mammalian G1 cyclin-dependent

kinases. Genes Dev., 9, 1149-1163.

215. Sherr,C.J. and Roberts,J.M. (1999). CDK inhibitors: positive and negative regulators of

G1-phase progression. Genes Dev., 13, 1501-1512.

216. Sherr,C.J. and Roberts,J.M. (2004). Living with or without cyclins and cyclin-dependent

kinases. Genes Dev., 18, 2699-2711.

217. Shih,I. and Kurman,R.J. (2002). Molecular basis of gestational trophoblastic diseases.

Curr. Mol. Med., 2, 1-12.

218. Shin,J.K., Jeong,Y.T., Jo,H.C., Kang,M.Y., Chang,I.S., Baek,J.C., Park,J.K., Lee,S.A.,

Lee,J.H., Choi,W.S., and Paik,W.Y. (2009). Increased interaction between heat shock

169

protein 27 and mitogen-activated protein kinase (p38 and extracellular signal-regulated

kinase) in pre-eclamptic placentas. J. Obstet. Gynaecol. Res., 35, 888-894.

219. Sibai,B., Dekker,G., and Kupferminc,M. (2005). Pre-eclampsia. Lancet, 365, 785-799.

220. Sibai,B.M. (2003). Diagnosis and management of gestational hypertension and

preeclampsia. Obstet. Gynecol., 102, 181-192.

221. Slim,R. and Mehio,A. (2007). The genetics of hydatidiform moles: new lights on an

ancient disease. Clin. Genet., 71, 25-34.

222. Slingerland,J. and Pagano,M. (2000). Regulation of the cdk inhibitor p27 and its

deregulation in cancer. J. Cell Physiol, 183, 10-17.

223. Smith,S.C., Baker,P.N., and Symonds,E.M. (1997). Placental apoptosis in normal human

pregnancy. Am. J. Obstet. Gynecol., 177, 57-65.

224. Smith,S.C., Leung,T.N., To,K.F., and Baker,P.N. (2000). Apoptosis is a rare event in

first-trimester placental tissue. Am. J. Obstet. Gynecol., 183, 697-699.

225. Smith,S.C., Price,E., Hewitt,M.J., Symonds,E.M., and Baker,P.N. (1998). Cellular

proliferation in the placenta in normal human pregnancy and pregnancy complicated by

intrauterine growth restriction. J. Soc. Gynecol. Investig., 5, 317-323.

226. Soleymanlou,N., Jurisica,I., Nevo,O., Ietta,F., Zhang,X., Zamudio,S., Post,M., and

Caniggia,I. (2005a). Molecular evidence of placental hypoxia in preeclampsia. J. Clin.

Endocrinol. Metab, 90, 4299-4308.

227. Soleymanlou,N., Jurisicova,A., Wu,Y., Chijiiwa,M., Ray,J.E., Detmar,J., Todros,T.,

Zamudio,S., Post,M., and Caniggia,I. (2007). Hypoxic Switch in Mitochondrial Myeloid

Cell Leukemia Factor-1/Mtd Apoptotic Rheostat Contributes to Human Trophoblast Cell

Death in Preeclampsia. Am. J. Pathol..

228. Soleymanlou,N., Wu,Y., Wang,J.X., Todros,T., Ietta,F., Jurisicova,A., Post,M., and

Caniggia,I. (2005b). A novel Mtd splice isoform is responsible for trophoblast cell death

in pre-eclampsia. Cell Death. Differ., 12, 441-452.

229. Soothill,P.W., Nicolaides,K.H., Rodeck,C.H., and Campbell,S. (1986). Effect of

gestational age on fetal and intervillous blood gas and acid-base values in human

pregnancy. Fetal Ther., 1, 168-175.

230. Soto-Wright,V., Bernstein,M., Goldstein,D.P., and Berkowitz,R.S. (1995). The changing

clinical presentation of complete molar pregnancy. Obstet. Gynecol., 86, 775-779.

231. St Jacques,S., Forte,M., Lye,S.J., and Letarte,M. (1994). Localization of endoglin, a

transforming growth factor-beta binding protein, and of CD44 and integrins in placenta

during the first trimester of pregnancy. Biol. Reprod., 51, 405-413.

170

232. Steigrad,S.J. (2003). Epidemiology of gestational trophoblastic diseases. Best. Pract. Res.

Clin. Obstet. Gynaecol., 17, 837-847.

233. Straszewski-Chavez,S.L., Abrahams,V.M., and Mor,G. (2005). The role of apoptosis in

the regulation of trophoblast survival and differentiation during pregnancy. Endocr. Rev.,

26, 877-897.

234. Susin,S.A., Zamzami,N., Castedo,M., Hirsch,T., Marchetti,P., Macho,A., Daugas,E.,

Geuskens,M., and Kroemer,G. (1996). Bcl-2 inhibits the mitochondrial release of an

apoptogenic protease. J. Exp. Med., 184, 1331-1341.

235. Suzuki,A., Tsutomi,Y., Akahane,K., Araki,T., and Miura,M. (1998). Resistance to Fas-

mediated apoptosis: activation of caspase 3 is regulated by cell cycle regulator p21WAF1

and IAP gene family ILP. Oncogene, 17, 931-939.

236. Szulman,A.E. and Surti,U. (1984). Complete and partial hydatidiform moles: cytogenetic

and morphological aspects. Adv. Exp. Med. Biol., 176, 135-146.

237. Tanaka,S., Kunath,T., Hadjantonakis,A.K., Nagy,A., and Rossant,J. (1998). Promotion of

trophoblast stem cell proliferation by FGF4. Science, 282, 2072-2075.

238. Tham,B.W., Everard,J.E., Tidy,J.A., Drew,D., and Hancock,B.W. (2003). Gestational

trophoblastic disease in the Asian population of Northern England and North Wales.

BJOG., 110, 555-559.

239. Timmer,J.C. and Salvesen,G.S. (2007). Caspase substrates. Cell Death. Differ., 14, 66-

72.

240. Vaisbuch,E., Ben Arie,A., Dgani,R., Perlman,S., Sokolovsky,N., and Hagay,Z. (2005).

Twin pregnancy consisting of a complete hydatidiform mole and co-existent fetus: report

of two cases and review of literature. Gynecol. Oncol., 98, 19-23.

241. Venkatesha,S., Toporsian,M., Lam,C., Hanai,J., Mammoto,T., Kim,Y.M., Bdolah,Y.,

Lim,K.H., Yuan,H.T., Libermann,T.A., Stillman,I.E., Roberts,D., D'Amore,P.A.,

Epstein,F.H., Sellke,F.W., Romero,R., Sukhatme,V.P., Letarte,M., and Karumanchi,S.A.

(2006). Soluble endoglin contributes to the pathogenesis of preeclampsia. Nat. Med., 12,

642-649.

242. Vervoorts,J. and Luscher,B. (2008). Post-translational regulation of the tumor suppressor

p27(KIP1). Cell Mol. Life Sci., 65, 3255-3264.

243. Villar,J., Carroli,G., Wojdyla,D., Abalos,E., Giordano,D., Ba'aqeel,H., Farnot,U.,

Bergsjo,P., Bakketeig,L., Lumbiganon,P., Campodonico,L., Al Mazrou,Y.,

Lindheimer,M., and Kramer,M. (2006). Preeclampsia, gestational hypertension and

intrauterine growth restriction, related or independent conditions? Am. J. Obstet.

Gynecol., 194, 921-931.

244. Vlach,J., Hennecke,S., and Amati,B. (1997). Phosphorylation-dependent degradation of

the cyclin-dependent kinase inhibitor p27. EMBO J., 16, 5334-5344.

171

245. Waga,S., Hannon,G.J., Beach,D., and Stillman,B. (1994). The p21 inhibitor of cyclin-

dependent kinases controls DNA replication by interaction with PCNA. Nature, 369,

574-578.

246. Wake,N., Seki,T., Fujita,H., Okubo,H., Sakai,K., Okuyama,K., Hayashi,H., Shiina,Y.,

Sato,H., Kuroda,M., and . (1984). Malignant potential of homozygous and heterozygous

complete moles. Cancer Res., 44, 1226-1230.

247. Watanabe,G., Albanese,C., Lee,R.J., Reutens,A., Vairo,G., Henglein,B., and Pestell,R.G.

(1998). Inhibition of cyclin D1 kinase activity is associated with E2F-mediated inhibition

of cyclin D1 promoter activity through E2F and Sp1. Mol. Cell Biol., 18, 3212-3222.

248. Weier,J.F., Weier,H.U., Jung,C.J., Gormley,M., Zhou,Y., Chu,L.W., Genbacev,O.,

Wright,A.A., and Fisher,S.J. (2005). Human cytotrophoblasts acquire aneuploidies as

they differentiate to an invasive phenotype. Dev. Biol., 279, 420-432.

249. Wilson,B.J., Watson,M.S., Prescott,G.J., Sunderland,S., Campbell,D.M., Hannaford,P.,

and Smith,W.C. (2003). Hypertensive diseases of pregnancy and risk of hypertension and

stroke in later life: results from cohort study. BMJ, 326, 845.

250. Wolter,K.G., Hsu,Y.T., Smith,C.L., Nechushtan,A., Xi,X.G., and Youle,R.J. (1997).

Movement of Bax from the cytosol to mitochondria during apoptosis. J. Cell Biol., 139,

1281-1292.

251. Wong,S.Y., Ngan,H.Y., Chan,C.C., and Cheung,A.N. (1999). Apoptosis in gestational

trophoblastic disease is correlated with clinical outcome and Bcl-2 expression but not

Bax expression. Mod. Pathol., 12, 1025-1033.

252. Woo,M., Hakem,A., Elia,A.J., Hakem,R., Duncan,G.S., Patterson,B.J., and Mak,T.W.

(1999). In vivo evidence that caspase-3 is required for Fas-mediated apoptosis of

hepatocytes. J. Immunol., 163, 4909-4916.

253. Woo,M., Hakem,R., Furlonger,C., Hakem,A., Duncan,G.S., Sasaki,T., Bouchard,D.,

Lu,L., Wu,G.E., Paige,C.J., and Mak,T.W. (2003). Caspase-3 regulates cell cycle in B

cells: a consequence of substrate specificity. Nat. Immunol., 4, 1016-1022.

254. Xue,W.C., Feng,H.C., Chan,K.Y., Chiu,P.M., Ngan,H.Y., Khoo,U.S., Tsao,S.W.,

Chan,K.W., and Cheung,A.N. (2005). Id helix-loop-helix proteins are differentially

expressed in gestational trophoblastic disease. Histopathology, 47, 303-309.

255. Yakovlev,A.G., Di Giovanni,S., Wang,G., Liu,W., Stoica,B., and Faden,A.I. (2004).

BOK and NOXA are essential mediators of p53-dependent apoptosis. J. Biol. Chem.,

279, 28367-28374.

256. Yang,J., Liu,X., Bhalla,K., Kim,C.N., Ibrado,A.M., Cai,J., Peng,T.I., Jones,D.P., and

Wang,X. (1997). Prevention of apoptosis by Bcl-2: release of cytochrome c from

mitochondria blocked. Science, 275, 1129-1132.

172

257. Yasuda,M., Umemura,S., Osamura,R.Y., Kenjo,T., and Tsutsumi,Y. (1995). Apoptotic

cells in the human endometrium and placental villi: pitfalls in applying the TUNEL

method. Arch. Histol. Cytol., 58, 185-190.

258. Yinon,Y., Nevo,O., Xu,J., Many,A., Rolfo,A., Todros,T., Post,M., and Caniggia,I.

(2008). Severe intrauterine growth restriction pregnancies have increased placental

endoglin levels: hypoxic regulation via transforming growth factor-beta 3. Am. J. Pathol.,

172, 77-85.

259. Zhong,X.Y., Laivuori,H., Livingston,J.C., Ylikorkala,O., Sibai,B.M., Holzgreve,W., and

Hahn,S. (2001). Elevation of both maternal and fetal extracellular circulating

deoxyribonucleic acid concentrations in the plasma of pregnant women with

preeclampsia. Am. J. Obstet. Gynecol., 184, 414-419.

260. Zhou,Y., Damsky,C.H., and Fisher,S.J. (1997). Preeclampsia is associated with failure of

human cytotrophoblasts to mimic a vascular adhesion phenotype. One cause of defective

endovascular invasion in this syndrome? J. Clin. Invest, 99, 2152-2164.

261. Zhou,Y., Genbacev,O., Damsky,C.H., and Fisher,S.J. (1998). Oxygen regulates human

cytotrophoblast differentiation and invasion: implications for endovascular invasion in

normal pregnancy and in pre-eclampsia. J. Reprod. Immunol., 39, 197-213.

262. Zinkel,S., Gross,A., and Yang,E. (2006). BCL2 family in DNA damage and cell cycle

control. Cell Death. Differ., 13, 1351-1359.

263. Zybina,T.G., Frank,H.G., Biesterfeld,S., and Kaufmann,P. (2004). Genome

multiplication of extravillous trophoblast cells in human placenta in the course of

differentiation and invasion into endometrium and myometrium. II. Mechanisms of

polyploidization. Tsitologiia, 46, 640-648.

264. Zybina,T.G., Kaufmann,P., Frank,H.G., Freed,J., Kadyrov,M., and Biesterfeld,S. (2002).

Genome multiplication of extravillous trophoblast cells in human placenta in the course

of differentiation and invasion into endometrium and myometrium. I. Dynamics of

polyploidization. Tsitologiia, 44, 1058-1067.

265. Zybina,T.G. and Zybina,E.V. (2005). Cell reproduction and genome multiplication in the

proliferative and invasive trophoblast cell populations of mammalian placenta. Cell Biol.

Int., 29, 1071-1083.