role of erk1/2 in the crosstalk between the pdgf- and...

122
Role of ERK1/2 in the Crosstalk between the PDGF- and Estrogen-Signalling Pathways in Neonatal Testicular Gonocyte Proliferation Monty Mazer Experimental Medicine McGill University, Montreal August 2010 A Thesis submitted to McGill University in partial fulfillment of the requirements of the degree of M.Sc. © Monty Mazer, 2010

Upload: ngoquynh

Post on 01-Apr-2019

216 views

Category:

Documents


0 download

TRANSCRIPT

Role of ERK1/2 in the Crosstalk between the PDGF- and Estrogen-Signalling Pathways in

Neonatal Testicular Gonocyte Proliferation

Monty Mazer

Experimental Medicine

McGill University, Montreal

August 2010

A Thesis submitted to McGill University in partial fulfillment of the requirements of the degree

of M.Sc.

© Monty Mazer, 2010

Abstract

Gonocytes are the precursors of spermatogonial stem cells from which spermatozoa

originate. We have shown that neonatal rat gonocytes proliferate in response to the combined

action of PDGF and 17β-estradiol (E2). The xenoestrogens Bisphenol A and genistein, previously

shown to alter the male reproductive system, stimulated gonocyte proliferation through

crosstalk with the PDGF pathway in a manner similar to E2, while testosterone and

progesterone did not affect gonocyte proliferation. Gonocytes expressed Raf1, MEK1, ERK1/2

and PI3K, and proliferated through ERK1/2 activation. PDGF and estrogen induced rapid ERK2

phosphorylation and their combination maintained activated ERK2 for 60 minutes, localized

mainly in the cytosol. E2 induced a rapid increase of estrogen receptor β immunoreactivity in

gonocyte cytosol. PDGF increased a cytosolic PDGFRβ signal, suggesting a role for the variant

V1-PDGFRβ previously identified in gonocytes. These data suggest that PDGF and estrogen are

required to maintain ERK2 activation which mediates gonocyte proliferation, and that estrogen

exerts rapid non-genomic effects in gonocytes.

Resumé

Les gonocytes sont les précurseurs des cellules souche spermatogoniales dont sont issus

les spermatozoa. Nous avons montré que les gonocytes néonataux de rat prolifèrent en

réponse à la combinaison de PDGF et d' estradiol (E2). Les xenoestrogenes Bisphenol A et

genistein, connus pour leurs effets perturbateurs sur le système reproductif masculin,

stimulaient la proliferation des gonocytes par un mécanisme similaire à celui de l' estradiol,

tandis que ni la testosterone ni la progesterone n' avaient d' effet sur la proliferation. Les

gonocytes exprimaient Raf1, MEK1, ERK1/2 et PI3K, et proliferaient via activation de ERk1/2.

PDGF et estrogen induisaient une rapide phosphorylation de ERK2, leur action combinée

maintenant cette phosphorylation pour 60 min, principalement dans le cytosol. E2 induisait

aussi l' augmentation rapide de l'immunoreactivité de l'estrogène receptor β dans le cytosol.

PDGf augmentait le signal cytosolic de PDGFRβ, suggerant un role du variant V1-PDGFRβ,

identifié préalablement dans les gonocytes. Ces résultats suggèrent que le PDGF et l'estrogène

sont nécessaires à la maintenance de l'activation de ERK2, médiateur de la prolifération, et que

l'estrogene exerce des effets rapides et non-génomiques dans les gonocytes.

Acknowledgments

I would like to graciously thank my Master’s supervisor, Dr. Martine Culty for all of her guidance and assistance with my research. Dr. Culty was instrumental in setting up my project and afforded me the tools to plan experiments independently while always providing insights and direction when necessary. Dr. Culty has been incredibly understanding, patient and helpful in providing me an excellent education during my years working in her laboratory, while ensuring that the research lab was a comfortable environment for everyone to be optimally productive. I am especially grateful to her for all the time that she spent helping me with papers, presentations and especially with writing this thesis, and as busy as she might have been, she always found time to meet to discuss results or to help me with anything I needed.

I would also like to acknowledge my coworkers, Annie Boisvert and Gurpreet Manku who have been incredibly helpful in assisting me throughout the course of my studies and in the process; we have also become very good friends. They were always there to help with anything I needed, and were reliable people to turn to for guidance when it was necessary. It was a true pleasure working with them.

As well, I have to thank the rest of the Culty / Papadopoulos labs for all their help and insights throughout. My two years working in the lab were truly enjoyable, and the people who worked with me created a friendly and warm working atmosphere.

The most important thank you must go to my endlessly supportive and helpful family, and especially to my loving wife Daniella, who stood by me through many late nights and weeks with countless hours of work as I balanced numerous things at once in order to finish my thesis. She showed her support every single day and was always understanding of the pressures of a Master’s student. No matter what, she was always there when I needed her and eager to help in any way possible to relieve some of the load. She definitely deserves a vacation as much as I do.

This thesis is dedicated in memory of my grandfather, Sid Mazer ל"ז , who passed away less than two weeks ago. He never stopped encouraging me to reach for the sky, and until the last day that I spoke with him was asking when I would finish my thesis. I have learned so much from him throughout my life, and his determination and his “never give up” attitude are lessons that I have always admired and continue to personify in everything I do.

Abbreviations

- 4OHT: 4-hydroxytamoxifen

- 3βAdiol: 5α Androstane-3β, 17β-diol

- As: Spermatogonia ASingle

- Apr: Spermatogonia Apaired

- Aal: Spermatogonia Aaligned

- AdDP: 4,4’-(1,3-adamantanediyl)diphenol

- AdP: Adamantly substituted phenol (4-(1-anamantyl)phenol)

- AdMP: 2-(1-adamantyl)-4-methylphenol

- AF-(1/2):activation factor 1/2

- AP-1: Activator protein 1

- APP: amyloid precursor peptide

- AR: androgen receptor

- ArKO: aromatase knock out

- BMP: Bone morphogenic protein

- BPA: Bisphenol-A

- cAMP: cyclic AMP (adenosine monophosphate)

- CIS: Carcinoma in situ

- CREB: cAMP response element binding protein

- CV: Cardiovascular

- CYP19: cytochrome P-450 aromatase

- Cyp40: cyclophilin 40

- DAG: 1,2-diacylglycerol

- DBD: DNA binding domain

- DES: Diethylstilbestrol

- DHT: 5α-dihydrotestosterone

- DNA: deoxyribonucleic acid

- dpc: Days post coitum

- E2: 17β-estradiol

- EGF: epidermal growth factor

- ER: estrogen receptor

- ERE: estrogen response element

- ERK: estracellular signal regulated protein kinase

- ERKO: estrogen receptor knock out mouse

- ERR: estrogen related receptor

- FBS: Foetal bovine serum

- FSH: Follicle-stimulating hormone

- GDNF: Glial cell derived neurotrophic factor

- GPCR: G-protein coupled receptor

- Grb2/7: Growth factor receptor bound protein 2/7

- GTPase: enzyme to hydrolyze guanosine triphosphate

- HGF: hepatocyte growth factor

- HSP: Heat shock protein

- ICC: Immunocytochemistry

- IHC: Immunohistochemistry

- IGF-1: Insulin-like growth factor 1

- IGF-1R: Insulin-like growth factor 1 receptor

- Il-1β: interleukin 1β

- IP3: inositol triphosphate

- JAK: Janus Kinase

- JNK: c-jun N-terminal kinase

- kDa: Kilodalton

- KSR: kinase suppressor of ras

- LBD: ligand binding domain

- LH: Luteinizing hormone

- MAPK: Mitogen activated protein kinase

- MAPKK: MAPK kinase (MEK)

- MAPKKK: MAPK kinase kinase (MEK kinase)

- MEK: MAPK/ERK kinase

- mER: membrane-bound estrogen receptor

- MKP: MAPK phosphatase

- MNAR: modulator of non-genomic activity of estrogen receptor

- MP1: MEK partner 1

- MTA1-S: metastatic tumour antigen 1 short form

- NES: nuclear export signal

- NF-κB: Nuclear factor- kappa B

- Ngn3: Neurogenin 3

- NR: nuclear receptor

- Oct-3/4: Octamer-4

- p75NTP: p75 neurotrophin marker

- PDGF: Platelet-derived growth factor

- PDGFR: PDGF receptor

- PDK1: pyruvate dehydrogenase kinase isoenzyme 1

- PDPN: Podoplanin

- PGC: Primordial germ cell

- PI3K: phosphatidylinositol 3-kinase

- PIP2: phosphatidylinositol-4,5-bisphosphate (PI 4,5 P2)

- PIP3: phosphatidylinositol-3,4,5-triphosphate (PtdIns (3,4,5) P3)

- PKA: Protein kinase A

- PKC: Protein kinase C

- PLC: Phospholipase C

- PMC: peritubular myoid cell

- PND: postnatal day

- PP2A: Protein phosphatase 2A

- ptch1: Protein patched homologue 1

- PTP-SL: Phosphotyrosine-specific phosphatase-SL

- RA: Retinoic acid

- Rap1: Ras proximate 1

- RAR-RXR: Retinoic acid receptor – Retinoid X receptor complex

- ROS: reactive oxygen species

- RTK: Receptor Tyrosine kinase

- SAPK: stress activated protein kinase

- SCF: Stem cell factor

- SERM: selective ER modulator

- SH2/3: Src homology 2/3

- SSC: Spermatogonial stem cell

- SOS: Son of sevenless

- Sry: Sex determining region Y gene

- STAT: Signal transducer and activator of transcription protein

- TDS: Testicular dysgenesis syndrome

- VEGF: vascular endothelial growth factor

- TFAP2C: Transcription factor AP-2 gamma

- TGCT: Testicular germ cell tumour

- TGF: transforming growth factor

- TNF-α: tumour necrosis factor α

- V1-PDGFRβ: Variant form of the PDGFRβ

Publications

Thuillier R, Mazer M, Manku G, Boisvert A, Wang Y, Culty, M (2010). Interdependence of platelet-derived growth factor and estrogen-signaling pathways in inducing neonatal rat testicular gonocytes proliferation. Biology of Reproduction, 82(5): 825-836

*A portion of the work for this thesis was published in this paper.

Table of Contents

1. Introduction 1

2. Germ Cells and Foetal Testis Development 1

2.1. Testis Structure 1

2.2. Germ Cell Origin 2

2.3. Foetal Testis Formation and Development 3

2.4. Neonatal Gonocyte Development 5

3. Spermatogenesis 9

3.1. Stem cell Renewal mechanisms 10

3.2. Hormonal Control of Spermatogenesis 13

4. The Study of Gonocytes

4.1. Scientific Models Appropriate for the Study of Neonatal Germ Cells 14

5. Testicular Dysgenesis Syndrome 15

6. Platelet Derived Growth Factor Signalling Pathway

6.1. PDGF signalling molecule 17

6.2. PDGF Receptors 18

6.3. PDGF in Testis Development and Function 21

6.4. Effect of PDGF on gonocytes 23

6.5. V1-Variant form of PDGFRβ 24

6.6. Pathologies Involving the PDGF Signalling Pathway 24

7. Extracellular-Stimulated Downstream Signalling Pathways 25

7.1. Mitogen Activated Protein Kinase Pathway 25

7.2. Phosphatidylinositol 3-Kinase 28

8. Estrogen

8.1. Endogenous Estrogens 29

8.2. Exogenous Estrogens 31

8.3. Estrogen Receptors

8.3.1. Genomic Function of the Estrogen Receptor 32

8.3.2. Estrogen Receptor Structure 33

8.3.3. Non-Genomic effects of ERs 34

8.4. Reproductive Effects of Estrogens

8.4.1. Estrogen signalling in Females 34

8.4.2. Estrogen Signalling in Male Reproductive System 35

8.5. Estrogen in Male Reproductive Development 36

8.6. The Estrogen Hypothesis 38

8.7. Approaches Used to Study the Role of Estrogen/ERs in Males

8.7.1. Laboratory results of estrogen exposure in vivo 40

8.7.2. Knockout Mice 42

9. Cell Signalling Cross Talk Mechanisms 45

9.1. Intercommunication of separate downstream pathways 46

9.2. Crosstalk between Estrogen Receptor and Growth

Factor Receptors / MAPK Pathway 47

10. Summary 51

11. Materials and Methods

11.1. Gonocyte Isolation 53

11.2. Cell Culture for Short Term Molecular Profile 56

11.3. PDGF-depleted FBS 57

11.4. Protein Analysis – Western Blot 57

11.5. Nuclear Isolation 59

11.6. Immunocytochemistry 59

11.7. Proliferation Assay 60

11.8. Immunohistochemistry 61

11.9. V1-PDGFRβ Vector Transfection and Live Cell Imaging 61

12. Results

12.1. Charcoal Stripped FBS and PDGF-Depleted Serum 63

12.2. Gonocyte Expression of Downstream Molecules

of the PDGF Signalling Pathway 63

12.3. In Vitro Exposure to Xenoestrogens and Phytoestrogens

Induce Proliferation in Neonatal Gonocytes 65

12.4. Treatment of Gonocytes with Other Steroid Hormones 68

12.5. ERK2 Activation via PDGF-BB and 17β-estradiol 72

12.6. PDGF and Estrogen Increase Expression of PDGFRβ

and ERβ Immunoreactivity In Vitro 77

12.7. Live Cell Imaging of Gonocytes Transfected with an

EGFP-V1-PDGFRβ Construct, Preliminarly Observations 78

13. Discussion 90

14. Conclusion 97

1. Introduction

Spermatogenesis is the male germ cell pathway necessary for procreation and

regeneration of the species. For viable fertilization to occur in any species, a high percentage of

healthy haploid gametes must be produced by the reproductive center of the organism. This

pathway in mammals is comprised of numerous cell divisions, regulatory mechanisms, positive

and negative feedback, and a host of other processes, all necessary for successful production of

fertile spermatozoa. While every step of germ cell progression between fertilization of a new

zygote and the organism’s subsequent production of sperm at the onset of puberty is crucial,

many of the intermediate stages are rarely discussed and not often investigated. Of particular

note, many scientists overlook the entire gonocyte stage. Gonocytes are cells that differentiate

from primordial germ cells and, through multiple mitotic phases and a lengthy quiescent

period, progress to spermatogonial stem cells, the first cells of the spermatogenetic pathway.

More detailed study of this area is necessary, given the evidence that the precursor to

Testicular Germ Cell Tumours (TGCT), Carcinoma in Situ (CIS), originates from this stage of

development.

2. Germ Cells and Foetal Testis Development

2.1 Testis Structure

The male reproductive system is made up of a number of organs which produce and

harbour the cells necessary for reproduction. The male testes, or gonads, are the organs where

spermatogenesis takes place. They are surrounded by a fibrous enveloping capsule and are

split into two separate compartments, the interstitium and the seminiferous tubules. The

interstitium is comprised of the Leydig cells as well as abundant vasculature, lymphatic vessels

and macrophages. The seminiferous tubules are a system of well organized convoluted tubules

which connects at the end to the rete testis. The seminiferous tubules are bound in place by

endothelial cells and are formed by an outer basement membrane covered with peritubular

myoid cells. The inside of the tubules consist of the germ cells surrounded by Sertoli cells.

Peritubular myoid cells (PMCs) are contractile cells which among other functions are used for

motility of the sperm through the tubules to the rete testis. Important for the growth and

1

hormonal regulation of germ cell development, the Leydig cells are the major source of the

androgen testosterone as well as many other hormones in the adult testis (Russell et al, 1990).

Sertoli cells are the somatic cells of the seminiferous tubules, bound to the germ cells by

intercellular gap junctions (Orth and Boehm, 1990) and they are responsible for providing

nutrition and structural support to the germ cells (Gnessi et al, 1995). Sertoli cells align their

nuclei along the basal membrane and extend their cytoplasm and extracellular matrices into

the middle of the tubule in immature testes, and retract to form a lumen in mature testes.

They function in regulating spermatogenesis, secrete liquid to fill the lumen and deliver

nutrition to the germ cells. They cease their mitotic divisions at puberty and Sertoli cell

numbers remain stable throughout the remainder of the man’s life.

2.2 Germ Cell Origin

Germ cells are the only cells of multicellular organisms that can undergo meiosis. They

are the cells that are capable of combining with a germ cell of the opposite sex to create new

organisms of a given species (Alberts et al, 2002). The germ cell lineage in both males and

females before gender specification and differentiation, begin as Primordial Germ Cells (PGCs),

pluripotent cells in the embryonal ectoderm, originating as part of the epiblast (Rouiller-Fabre

et al, 2003). Germ cells are first seen in mice at 7.5 days post coitum (dpc) and are

characterized by their alkaline phosphatase activity, as well as their retention of the

transcription factor Oct-4, a protein originally expressed in all totipotent embryonic cells that

becomes restricted to PGC. They total nearly 100 cells (Ohmura et al, 2004;De Rooij,

1998;Ohbo et al, 2003). It is of interest that there is no direct germ cell lineage from

fertilization. The germ cells originate as normal pluripotent embryonal cells that are induced to

PGC specification under the control of the bone morphogenic proteins BMP-4 and 8b at their

specific location in the extraembryonic region posterior to the primitive streak (Ying et al,

2000;Lawson et al, 1999). This group of cells migrates through the primitive streak towards the

exoderm where some cells remain in the genital ridge as primordial germ cells, while others

continue onwards to form the mesoderm (McLaren, 1999). PGC migration is dependent on

their expression of the receptor c-KIT (Sutton, 2000) a protein also found in a number of

2

embryonic stem cells (Ashman et al, 1991). PGC migration in humans begins at the stalk of the

allantois during the 4th week of gestation until week 5 when they reach the gonad. Testis cord

formation is then initiated over the next two weeks until week 7 (Lambrot et al, 2006). At this

point, the genital ridge, or gonad, is bipotent and can develop into either an ovary or testis. Sex

determination is decided based on the expression of the Sry gene on the Y chromosome in the

somatic cells destined to become foetal Sertoli cells. The expression of the Sry gene begins on

gestational day 10.5 and continues until day 12.5, inducing expression of the transcription

factor Sox9. The expression of both these factors is essential for development of the male

gonad, and failure to express either Sry or Sox9 will default the gonad into forming an ovary

(Basciani et al, 2010). Epigenetic changes as well as the majority of genetic imprinting in the

germ cell lineage takes place around 10.5 dpc. More imprints are set in place prior to birth

(Chuma et al, 2005). The testis begins to be formed on 11.5 dpc in mice and 12.5 dpc in rats

when the PGCs arrive at the genital ridge and are surrounded by Sertoli cells, which are

differentiating from their embryological precursor into the gonadal somatic cells, thereby

creating the seminiferous tubules (Lambrot et al, 2006).

2.3 Foetal Testis Formation and Development

Testis formation requires a tightly regulated and coordinated sequence of proliferation,

migration, apoptosis and differentiation affecting several somatic cell types and germ cells

(Puglianiello et al, 2004). Any changes or abnormalities to this process can lead to infertility or

cancer (Wang and Culty, 2007). By 13 dpc there are approximately 10,000 PGCs in each of the

two forming gonads (Chuma et al, 2005;De Rooij, 1998). Once embedded in the Sertoli cell

matrix, the PGCs are designated as gonocytes, although some authors refer to them as

prespermatogonia, prospermatogonia or postmigratory PGCs (Olaso and Habert, 2000).

Although there are not many morphological differences between PGCs and gonocytes, which

appear so far to present similar gene expression profiles (Culty, 2009;Gaskell et al, 2004) one

functional difference is that gonocytes can only be cultured in vitro in the presence of Sertoli

cells, while PGCs can be cultured with any somatic cell type (De Rooij, 1998). Germinative cells

are generally identified by high levels of alkaline phosphatase activity (Puglianiello et al, 2004).

3

Mesenchymal cells migrate from the mesonephros into the gonad and differentiate into

myoid cells, pericytes and endothelium, which are critical steps in organized testis formation

(Puglianiello et al, 2004). Foetal Leydig cells appear in the interstitium on 12.5 dpc (Schmahl et

al, 2008) and begin to mature starting at 14.5 dpc with the production of testosterone, which

contributes to the development of male sex characteristics. Although mature adult Leydig cells

are stimulated by luteinizing hormone (LH) to produce testosterone, at this point until

approximately 20 dpc, Leydig cells function independent of gonadotropins. There are two

different Leydig cell populations found in the testis interstitial tissue at different developmental

periods. Although both cell types serve similar roles as the main steroidogenic cells of the

testis, they are morphologically and functionally quite different, supporting the idea that these

two types of Leydig cells arise from dissimilar precursor cells lineages. While foetal Leydig cells

reach their functional peak around gestational day 19, adult Leydig cells become fully matured

by 56 days after birth in rodents. Foetal Ledig cells are smaller and more sporadically located in

the interstitium and begin to atrophy and disperse over the first two weeks after birth. At this

point, Leydig stem cells begin to develop and produce progenitor Leydig cells which are spindle

shaped cells expressing a receptor for LH as well as steroidogenic enzyme activity, though they

produce very little testosterone. The cells enlarge and decrease their proliferative abilities until

day 56, when they are immature Leydig cells producing mostly 5α-reduced androgens. A final

proliferation and differentiation step results in mature adult Leydig cells producing testosterone

for testis function (Dong et al, 2007).

Gonocytes undergo two active periods of proliferation separated by a quiescent period

spanning from 17.5 dpc until neonatal day 3. Although in rats and mice the second proliferative

stage takes place postnatally, in humans, gonocyte proliferation and migration all occur in the

gestational period and spermatogonia remain quiescent from birth until pre-puberty. During

the first and second active periods, gonocytes proliferate and simultaneously undergo

apoptosis (Lambrot et al, 2006). Indeed germ cell apoptosis occurs mainly after the first

postnatal week and during the second week, when the cells have differentiated into

spermatogonia (Jahnukainen et al, 2004). Cells that failed to migrate and become

spermatogonia are then eliminated by apoptosis (Tres and Kierszenbaum, 2005). Proliferation

4

in rat gonocytes continues until day 6 when the first spermatogonia can be identified (Boulogne

et al, 2003). Proliferation and apoptosis can regulate the Sertoli cell / gonocyte ratio.

Alternatively, it is proposed that the importance of building this germinitive pool through both

proliferation and apoptosis is to negatively select the abnormally developed gonocytes to

prevent serious fertility problems or defects in future offspring (Olaso and Habert, 2000).

Sertoli cells continue to grow until the 3rd postnatal week, enlarging the diameter and length of

the seminiferous cords (Boulogne et al, 2003).

2.4 Neonatal Gonocyte Development

Beginning on gestational day 17.5 following the first proliferative stage, gonocytes arrest

their mitotic cell cycle at the G0/G1 phase until after birth. Termed “reproliferation”,

gonocytes end their quiescent period and activate proliferation on neonatal day 3 in rats and

around day 1.5 in mice. Simultaneously, gonocytes migrate to the basement membrane in a

process identified by some as prespermatogenesis. To be consistent with their function,

Hilscher uses different nomenclature to describe gonocytes. He coined the terms ‘multiplying’

and ‘transitional-prospermatogonia’ (Hilscher, 1991). Independent of all factors outside of the

testis, gonocytes will continue to proliferate and migrate in vitro in organ culture or coculture

with Sertoli cells. It was first hypothesized by McGuinness and Orth (McGuinness and Orth,

1992), and subsequently proven by studies in many different species and strains of mice, that

migration occurs independently of proliferation, and both can take place independent of the

other process. Studies showed that although most gonocytes proliferated before they

migrated, there were also cells which migrated to the basement membrane prior to

proliferation. Nagano et al. showed that gonocyte migration began in mice on day 18.5 post

coitum and continued without proliferation until neonatal day 1.5. From then on, migrated

cells were seen to proliferate prior to, as well as after migration. These two major events in

prespermatogenesis must be regulated by completely different mechanisms. In order to

migrate, gonocytes extend pseudopods from their cytoplasm to move around the Sertoli cell

matrix. By neonatal day 5 or 6 in rats, all normal gonocytes have reached the basement

membrane (Nagano et al, 2000). Any abnormal gonocytes or poorly formed spermatogonia

5

that either had problems with migration or with other functions will ultimately degenerate and

be eliminated by apoptosis, which is seen at low levels in rat germ cells until postnatal day

(PND) 20 (Basciani et al, 2008;Roosen-Runge and Leik, 1968). As part of this systematic process,

following the proliferation and migratory stages, gonocytes begin to differentiate into more

mature germ cells (Boulogne et al, 2003).

The next logical step in the germ cell lineage is formation of spermatogonial stem cells

by gonocytes, which will then differentiate to give rise to type A spermatogonia as the first

official step of spermatogenesis at the onset of puberty. Although this appears the reasonable

pathway, it is most likely the case that the gonocytes are in fact a heterogeneous population

where a portion of cells are already committed in foetal or early neonatal life to differentiate

directly into spermatogonia of the first spermatogenic wave and not the adult stem cell type.

This fact was supported by the work of Yoshida et al. (2006) who showed a subset of neonatal

gonocytes led to the formation of differentiating spermatogonia negative for the transcription

factor neurogenin 3 (Ngn3), rather than to the generation of Ngn3-positive spermatogonial

stem cells. Although there are currently no specific markers that designate each discreet

subset of gonocytes, the simultaneous investigation of several gene sets has clearly shown that

gonocytes are not all uniform in their profiles and some gonocytes might express proteins more

similar to stem cells, while other are closer to spermatogonia (Culty, 2009;Yoshida et al,

2006;De Rooij, 1998).

In order to investigate the foetal germ cell commitment to spermatogenesis and the

ability for germ cells to populate an infertile environment, three important studies were

conducted that would eventually help mould our understanding of germ cell development and

function. Brinster and Zimmermann showed that transplantation of spermatogonial stem cells

and spermatogonia populations into an infertile mouse induced normal spermatogenesis and

fertile gametes (1994). In another study, Ohta et al. (2004) transplanted foetal gonocytes into

infertile mice to explore when gonocytes become committed to spermatogenesis. They

observed that although the 14.5 dpc gonocytes did produce viable spermatozoa, germ cells

from 12.5 dpc mice did not induce spermatogenesis. This time period correlates well with the

6

Mig

rato

ry (m

ouse

)

Prim

ordi

al G

erm

Cel

lG

onoc

yte/

Pres

perm

atog

onia

SSC/

Sper

mat

ogon

ia

Mou

se(d

ays)

Rat

(day

s)

Hum

an(w

eeks

)

7.5

13.5

1619

813

1821

1.5

Prem

igra

tory

Post

mig

rato

ry

11.5

37

5

Mito

ticQ

uies

cent B I R T H

3

4-5 5

Mito

tic

Mig

rato

ry

(rat

)

Apop

tosis

Apop

tosis

Cord

form

atio

n

820

A s>A

pr>

A al>

A1-

4>

In >

B 10 10

36 Birt

h

Sper

mat

ogen

esis

Gen

erat

ion

of sp

erm

atog

onia

l ste

m ce

lls

TGCT

sCI

S ce

lls

Mito

tic16

8-12

SSC

A pal

e/A d

ark/

B

Qui

esce

nt

P U B E R T Y

30-4

0

11-1

3 ye

ars

7

Figure 1: Early Male Germ cell Timeline. Timeline depicting the majorprocesses in the prespermatogenesis phase of male germ cell developmentuntil the onset of puberty in the human, rat and mouse. The first stagepresented in the diagram is the undifferentiated embryonic primordialgerm cell (PGC), and the last stage is the differentiated type Bspermatogonia. The picture under the timeline, beginning from the left,represents a cross section of a foetal seminiferous cord as it develops intoa seminiferous tubule. PGCs and gonocytes are found in the center of thecord, gonocytes migrate out to the basement membrane and the Sertolicell cytoplasm eventually retracts forming the lumen of the tubule. Onceadjacent to the basement membrane, gonocytes will differentiate intospermatogonial stem cells which will eventually produce sperm whichenters the lumen of the tubules. In the event of an error in one of theregulation mechanisms, gonocytes may not properly migrate anddifferentiate and may fail to be eliminated by apoptosis, leading tocarcinoma in situ (CIS), which can progress to form testicular germ celltumours (TGCT).

8

first proliferation stage of foetal gonocytes, and highlighted a critical functional difference

between PGC and their descendents the gonocytes. It is important to realize though, that germ

cells will only continue to differentiate and eventually undergo meiosis when in direct

communication with Sertoli cells. Finally, Chuma et al. transplanted epiblast cells and PGCs into

infertile mice, and showed that they were able to continue normal development in the

presence of mature Sertoli and Leydig cells to eventually become spermatogonial stem cells

and produce viable spermatozoa (Chuma et al, 2005).

3. Spermatogenesis

There are three major steps in spermatogenesis; proliferation of the germ cells,

separation of genetic material via meiosis, and spermatozoa development (Russell et al, 1990).

There are many different models for how spermatogenesis in rodents occurs, most of which are

quite similar, most often differing in where the germ cells irreversibly differentiate downstream

in the pathway. I will discuss briefly the Huckins and Oakberg’s (Huckins, 1971;Oakberg, 1971)

As model. Gonocyte differentiation, whether or not through extra mitotic divisions, results in

type A spermatogonia at the basement membrane of the seminiferous tubules. These first

spermatogonia are called Asingle (As) and function as the spermatogonial stem cell (SSC). SSCs

can divide to replenish the SSC population by creating two separate cells, which both act as

stem cells, or they can divide into two daughter cells that remain attached together through

intercellular bridges to become Apair (Apr) (De Rooij, 1998). Alternatively, it is possible that the

SSC will differentiate asymmetrically into one stem cell and another cell that will immediately

progress to spermatogenesis (De Rooij, 2001). Under normal conditions, SSCs renew

themselves at a 1:1 ratio with Apr cells. The Apr cells divide into 4, 8, 16 and rarely 32 Aaligned (Aal)

cells. Until this point, the cells are considered undifferentiated spermatogonia (De Rooij, 1998).

The first differentiation occurs as the Aal cells become A1, a process that is regulated by the

active metabolite of vitamin A, retinoic acid (De Rooij, 2001), as well as cyclin D2 (Beumer et al,

2000). The A1 cells subsequently divide six times, becoming A2, A3, A4, A-intermediate cells

and spermatogonia B (De Rooij, 1998). The next division defines the transformation to primary

spermatocytes, called preleptotene. Before meiosis can take place, the cells remain in

9

prophase for three weeks, during which they go through many phases, characterized by

increases in cellular and nuclear size as well as changes in chromatin conformation in

preparation for division. The designations given to the different phases are leptotene,

zygotene, pachytene and diplotene. The first division, meiosis I, creates two secondary

spermatocytes, and meiosis II splits these cells into spermatids (Russell et al, 1990). The final

process, which contains nineteen steps in rats and sixteen in mice, is called spermiogenesis.

The spermatids develop a flagellum, concentrate their nuclear material in the head which is

surrounded by an acrosome, and remove the remaining cytoplasm, creating the spermatozoa.

The spermatozoa are then released into the lumen of the tubules and travel to the rete testis.

They are stored in the epididymis for final maturation (chapter 6). Unlike in the rodent, human

spermatogenesis undergoes fewer divisions and only has one intermediate stage between the

SSC and type B spermatogonia, termed Apale Spermatogonia. In total, a single mouse SSC will go

through approximately 13 divisions yielding 8192 sperm, while human SSCs only go through 4

divisions and therefore yield only 16 sperm (Schlatt, 2010).

3.1 Stem Cell Renewal Mechanisms

Under normal circumstances, the 1:1 ratio of stem cells to Apr spermatogonia is

sufficient to maintain a large stem cell pool. In a case where the stem cell pool had been

depleted due to toxic substances or irradiation, the remaining stem cells are able to begin

renewing themselves at a higher than normal rate until the stem cell pool is returned to

normal. SSCs have been shown to preferentially occupy specific areas of the tubule periphery

which are in proximity of the blood vessels, defining a stem cell “niche” (Yoshida et al, 2007).

The stem cells occupy open areas on the Sertoli matrix, and when there are low numbers of

stem cells, Sertoli cells secrete high amounts of GDNF to stimulate stem cell renewal. It is

generally accepted that spermatogonia Apr and Aal are undifferentiated. The question arises

whether or not they maintain their stem cell capabilities. It has been noted that in some

mutant mice, there were some aligned spermatogonia with odd numbers of cells instead of

even numbers. It is conceivable that this is an emergency mechanism to replenish the stem cell

10

Spermatogonia

Dark

Pale B

Prim

ary

Sper

mat

ocyt

e (4

n)

Seco

ndar

y Sp

erm

atoc

yte

(2n)

Mei

osis

1 Mei

osis

2

SpermatogoniaA s A p

r

A al

A 1 A 2 A 3 A 4

B

Prim

ary

Sper

mat

ocyt

e (4

n)

Seco

ndar

y Sp

erm

atoc

yte

(2n)

Mei

osis

1

Mei

osis

2

Sper

mat

id(1

n)

Sper

mat

ozoa

Undifferentiated

DifferentiatingIn

term

edia

te

11

Figure 2: Spermatogenesis. Representative diagram of thespermatogenic cycle in human and rodents. The diagram shows themain steps of male germ cell development starting at thespermatogonial stem cell that take place throughout life. The cycleincludes 3 phases, the proliferative phase (spermatogonia), meioticphase (spermatocytes) and spermiogenesis (spermatidmetamorphose into spermatozoa). The first spermatogonial stepsare initiated before puberty in rodents. The whole cycle takesaround 50 days in rat and 64 days in human, leading to theformation of 100-200 millions spermatozoa in men and billions inrodents.

12

pool. Once the aligned cells differentiate to A1, the cells are then irreversibly directed towards

becoming spermatozoa (De Rooij, 2001).

3.2 Hormonal Control of Spermatogenesis

Testis function and embryonic development are regulated by the pituitary gonadotropic

hormones luteinizing hormone (LH) and follicle-stimulating hormone (FSH). Both FSH and LH

are secreted in cycles, a number of times each day, followed by a surge of testosterone. LH

stimulates Leydig cell testosterone production, which is secreted in the testes to regulate

spermatogenesis. Testosterone and FSH act indirectly on the germ cells though the Sertoli cells

to trigger spermatogenetic processes. FSH stimulates the Sertoli cells to secrete androgen

binding protein and inhibin which both act directly on the germ cells. All of the necessary

hypothalamic hormones are regulated negatively by testosterone levels (Strauss and Barbieri,

2004). Although the pituitary gonadotropins oversee complete testes development and

function, they cannot precisely regulate the detailed activities of individual cells in all of these

processes. For precise and accurate organization of the developing testis, there is a host of

paracrine and autocrine factors acting on the various cells (Gnessi et al, 1995). Included in

these factors are hepatocyte growth factors (HGFs), transforming growth factors (TGFs),

neurotrophins, and platelet-derived growth factors (PDGFs), which will be discussed later in

depth. HGF is critical for seminiferous cord formation, activating cell proliferation, and

migration (Ricci et al, 2004). Studies have shown that Retinoid acid (RA) has a small

proliferative effect on organ cultures of PND3 rat gonocytes (Livera et al, 2000) and in

gonocyte-Sertoli cell co-cultures (Boulogne et al, 2003), as well as in PND2 mice (Zhou et al,

2008a). However, RA treatment was shown to have the opposite effect on foetal and neonatal

gonocytes co-cultured for 3 to 6 days with a mixture of testicular somatic cells (Boulogne et al,

2003). It appears that retinoids, acting via specific RAR-RXR receptor dimers, affect both

mitosis and apoptosis in neonatal germ cells and these effects differ in function depending on

the conditions and cell types present in the cultures. Alternatively, Retinoic acid plays an active

role in gonocyte differentiation and maturation at the neonatal day 3 active stage in rat (Wang

and Culty, 2007) and in mouse (Zhou et al, 2008a;Zhou et al, 2008b).

13

4. The Study of Gonocytes

4.1 Scientific Models Appropriate for the Study of Neonatal Germ Cells

Investigation of germ cells in general, and more specifically, PGCs, Gonocytes and SSCs,

is extremely difficult for a number of reasons. PGCs and gonocytes are quite fragile in foetal life

and are not easily isolated nor identified due to lack of adequate specific markers. Studying

spermatogonial stem cells has the added difficulty of being one of a host of morphologically

similar cells in the testis. In 1994, Brinster’s lab discovered a functional test for SSCs, but low

cell numbers impairs the ability to study them. In addition, the study of most germ cell

functions require primary culture, as immortalizing germ cells would certainly change their

ability to proliferate and differentiate compared to in vivo (De Rooij, 2001). It is therefore

imperative that there be many different methods of gonocyte study, each of which examining

gonocyte function from a unique angle, enabling researchers to extract invaluable information

about a specific dimension about gonocytes. Of course, none of the methods are perfect, and

all of them must be employed in order to fully understand the big picture of foetal reproductive

development. In vivo studies commonly use transgenic mice or knockout genes to observe

testicular development following manipulation of specific factors or receptors. Studying murine

embryonic exposure to external factors such as estrogenic compounds or various

environmental toxins can also be useful in understanding early reproductive development.

Methods have been devised to study different actions of germ cells in vitro. Feeder cultures

and organ cultures are often used, which allow the tissue to continue its normal foetal and

neonatal growth in medium without serum or external nutrients (Lambrot et al, 2006;Livera et

al, 2006;Olaso and Habert, 2000). Organ culture, or organotypic culture, was developed to

create an in vivo-like environment for the germ cells in vitro (Rouiller-Fabre et al, 2003).

Gonocytes can also be purified and then cocultured with Sertoli cells or grown in medium alone

in order to dilute the testicular hormonal factors which can affect the development of these

cells. Long term gonocyte cultures require somatic cells in order for them to properly survive.

Using organ culture methods or cocultures with different somatic cells has a serious

shortcoming in that one cannot study the direct effects of a treatment of the germ cells without

14

accounting for an indirect effect of the treatment via somatic cells as well as the paracrine

relationship between germ cells and the surrounding somatic cells. Any contact with other cells

means that treatments will not only affect the germ cells, but will cause the somatic cells to

play a role in the reaction as well. Somatic cells are known to regulate germ cell growth by

secreted factors or directly through gap junctions. Alternatively, culturing germ cells alone

means that they are not in their native environment and might not function exactly as they

would in vivo. It is therefore crucial that one carefully decides the purpose of a study, before

choosing a method (Olaso and Habert, 2000).

5. Testicular Dysgenesis Syndrome

Over the past half century, there has been a steady increase in male reproductive

system disorders. These disorders include hypospadia, cryptorchidism, impaired

spermatogenesis, infertility and testicular cancer, all of which fall under the broader title of

testicular dysgenesis syndrome (TDS) (Sharpe and Skakkebaek, 2008). The syndrome

encompassing all of these developmental disorders describes how reproductive disorders are

all interconnected and ultimately can lead to infertility and testicular cancer. There is a direct

correlation between the number and severity of minor disorders and the development of some

type of reproductive cancer. The outcomes of TDS range from very few problems and almost

flawless fertility, to more severe developmental or pathogenic problems. Although there are

genetic abnormalities that can explain many of these disorders, it is generally accepted that the

main cause of the majority of TDS cases is environmentally related. There is a clear correlation

between the risks of TDS and the geographical location and environmental exposures of

pregnant mothers, suggesting that even testicular dysgenesis that is not apparent or that is

asymptomatic until much later in life, very often is predisposed by events during the life of the

developing foetus (Sonne et al, 2008). One of the main proposed causes of TDS is in utero

estrogen exposure which will be discussed in detail in a later section.

The incidence of testicular cancer has doubled since 1960, and germ cell testicular

cancer is the most common cancer in men aged 15 to 35 (Holmes, Jr. et al, 2008). Germ cell

tumours are almost completely curable if detected early and generally have a very positive

15

prognosis, although consequences of intense chemotherapy and radiotherapy can be harmful.

Effects include secondary malignancy and reduced fertility. If detected too late, testicular germ

cell tumours (TGCTs) can be lethal (Hoei-Hansen et al, 2007). Almost all germ cell tumours,

whether they develop into seminomas, or the more aggressive non-seminomas and embryomal

carcinomas, originate as Carcinoma in Situ (CIS), as defined by Skakkebaek in 1972 (Joensen et

al, 2007). Seminomas appear from primordial germ cells or gonocytes, and non-seminomas are

made up of neoplastic tissue, which usually consists of somatic or embryonal tissue (McIntyre

et al, 2005). The non-seminomas contain pluripotent cells which make up a heterogeneous

cancer population. Embryonal carcinomas can differentiate into different types of tissues,

producing a teratoma (Sonne et al, 2008). CIS can remain latent and symptom free for five to

fifteen years before becoming invasive (Hoei-Hansen et al, 2007). CIS cells are bigger than

spermatogonia and have a very large nucleus and well-defined nucleolus reminiscent of

gonocytes (Joensen et al, 2007). CIS testicular tissue has smaller and less developed

seminiferous tubules along with impaired spermatogenesis. It is very clear that CIS contains

distinct morphological similarities to primordial germ cells and gonocytes, and are believed to

originate directly from these lines of spermatogonial precursors in utero (Rajpert-De Meyts E.

and Hoei-Hansen, 2007;Sonne et al, 2008). These CIS cells which are derived from gonocytes or

their predecessor primordial germ cells and spermatogonial precursor cells, are inhibited from

differentiation and therefore always maintain immature germ cell morphology (Horwich et al,

2006). Not only do they have common morphological similarities to foetal/neonatal germ cells,

they have practically identical protein profiles, staining positive for stem cell markers such as c-

KIT, Oct3/4, NANOG, PDPN and TFAP2C (Joensen et al, 2007). Usually CIS cells remain in the

seminiferous tubules until puberty when they start to proliferate. These cells are able to enter

the lumen and can often be detected in the semen (Hoei-Hansen et al, 2007). With the onset of

pubertal proliferation, CIS cells undergo mitosis without differentiation, accumulating further

genetic mutations, en route towards genetic instability. This can eventually lead to invasive

testicular cancer (Sonne et al, 2008).

Given the prevalence of germ cell cancers and the increasing evidence of their gonocyte

origin, there is a growing need for investigative study in the field of foetal and neonatal

16

gonocyte growth and development. Unfortunately, the principal factor impeding CIS research

is that rodents do not appear to express a CIS or TGCT phenotype. For this reason, other

species and cell lines are necessary to complement rat and mouse studies. Nevertheless,

before experimenting on larger animals with more similar developmental patterns as humans,

one must gain a profound understanding of testis development in established model systems.

Ultimately, understanding normal testis developmental progression and mechanisms of

differentiation can lead to a better comprehension of what may go wrong to cause testicular

cancer (Joensen et al, 2007).

6. Platelet Derived Growth Factor Signalling Pathway

6.1 PDGF Signalling Molecule

Platelet derived growth factor (PDGF) is a growth factor that was first discovered in

blood plasma and, as the name suggests, was assumed to have originated in platelets. PDGF

was discovered as a molecule secreted into the plasma that could stimulate the proliferative

effect of fibroblasts. It is now known that PDGF is expressed in a large number of different

tissues and plays a role in many important activities in development (Basciani et al, 2010). The

paracrine activity of growth factors is a critical step in the communication and organization of

developing tissues (Ricci et al, 2004). Until roughly twenty years ago, there were only two

known PDGF isoforms; A and B. More recently C and D isoforms have been identified. Each

PDGF molecule, although expressing various degrees of homology, is transcribed from a unique

gene, on chromosomes 7, 22, 4 and 11 respectively in human. All of the PDGF hormones

contain six exons except for PDGF-D which is coded by seven. PDGF molecules are active in

their dimerized form, and dimerize, immediately after secretion, into homodimers or

heterodimers depending on the degree of expression of each of the molecules in a given tissue.

The only two PDGFs that are known to form heterodimers thus far are the A and B homologues.

Each PDGF contains eight cysteine residues in its polypeptide chain, forming the cysteine knot

which is responsible for intermolecular binding and dimerization. PDGF molecules have specific

sequence homology to another family of growth factors called the vascular endothelial growth

factor (VEGF) family. These two groups of growth factors share a conserved 80 – 90 amino acid

17

sequence which codes for the growth receptor binding domain. The propeptides of PDGF-A

and B are activated by proteolytic cleavage in the endoplasmic reticulum, while the C and D

isoform propeptides are only activated extracellularly. PDGF acts directly on specific receptors

to induce a wide variety of cellular functions such as proliferation, survival, angiogenesis,

migration, cytoskeletal rearrangements, as well as glycosaminoglycan, proteoglycan and

collagen production and secretion. It is also a promoter of tissue remodelling and embryonic

development of the kidneys, brain, lungs, heart and testis. Generally, PDGF molecules act upon

fibroblasts, neurons, endothelial cells and epithelium (Basciani et al, 2010).

6.2 PDGF Receptors

There are two specific PDGF receptors, alpha and beta, which are encoded on human

chromosomes four and five. The PDGF receptor (PDGFR) is a tyrosine kinase receptor

containing five extracellular immunoglobulin repeats, a single transmembrane and

juxtamembrane domain, two tyrosine kinase domains and a C-terminal domain. Binding PDGF

dimers causes receptor dimerization, and each receptor dimer has a unique affinity for specific

ligand isoforms. The αα homodimer binds to PDGF-AA, BB, AB and CC with preferential affinity

for AA and AB, the ββ homodimer binds to PDGF-BB and DD with highest affinity for BB, and the

αβ heterodimer binds to PDGF-AB, BB, CC and DD (Mariani et al, 2002). Upon ligand-binding,

the receptors are activated and autophosphorylated on tyrosine residues opening docking sites

for Src homology 2 (Sh2) and Sh3 domains. Activated PDGFRs can bind to Grb2, Shc, Grb7, Nck

and Crk which activate effectors such as PI3 kinase (PI3K), phospholipase C γ, STATs, Ras and

JAK/STAT. These effectors will trigger various downstream pathways of second messengers

such as; 1,2-diacylglycerol, inositol triphosphate, GTPase and MAP kinase, and downstream

transcription factors such as c-jun and c-fos. Generally, different receptors and ligands are

found to affect different tissues, and often different receptor/ligand patterns can generate

contradictory actions (Basciani et al, 2010). In cases where both types of PDGF receptors in a

single tissue or cell induce opposing actions on cell activity, it is very likely that the α-β

heterodimer will be involved in mediating the response (Mariani et al, 2002). The study of

PDGF receptor knockout mice have shown that PDGF-A is important for the development of

18

Grb

2

PV1

-PDG

FRβ

Ras

Raf

MEK

ERK

Surv

ival

Prol

ifera

tion

PP

PP

PP

PDG

FRβ/

βPD

GFR

α/α

DB

B B

A A

A D

C C

PDG

FRα/

β

C- Jun

C- Fos

JAK

P STAT

PP

STAT

STAT

PP

STAT

PI3K

AKT

Angi

ogen

esis

Mig

ratio

n

Colla

gen

and

GAG

prod

uctio

n

19

Figure 3: PDGF / PDGFR, receptor - ligand induced pathways. Representativediagram depicting the preferential affinity binding of PDGF A, B, C and D to thetwo receptors, alpha and beta. Three of the most common and well knownpathways are described in the schematic, all of which can be, and are found tobe, induced by each of the three PDGFR dimers. JAK/STAT: JAK is recruited toactivated autophosphorylated tyrosine kinase sites on the PDGFR and isphosphorylated itself, changing into its active form. It induces phosphorylationof STAT which dimerizes and enters the nucleus. STAT acts as a transcriptionfactor to promote transcription of genes. PI3K: Activated PDGFR canphosphorylate PI3K and activate it which will then travel downstream, andactivates a host of different molecules, most notably AKT (PKB). MAPK: PDGFRrecruits a Grb-SOS complex and activates it which will localize Ras to the plamamembrane in close proximity to the receptor. Activated Ras will then activatethe classic MAPK downstream pathway of Raf, MEK and ERK. ERK cantranslocate into the nucleus and modulate transcription.Gonocytes have been shown to express a variant form of PDGFRβ called V1-PDGFRβ. Its exact function is not yet known.

20

alveolar smooth muscle and oligodendrocytes, while PDGF-B is necessary for microvascular

pericytes, and kidney messangial cell migration and organization (Basciani et al, 2002). PDGF

function is regulated in the cells via specific mechanisms. SHP-2 binds to a phosphorylated

tyrosine near the C-terminal of PDGFRs and acts as a phosphatase, causing dephosphorylation

of the receptor (Lu et al, 1998). Activated PDGFRs are also often internalized into the cell and

degraded in lysosomes (Dai, 2010).

6.3 PDGF in Testis Development and Function

From early in embryogenesis, beginning at 12.5 dpc, Sertoli cells produce and secrete

PDGF in higher doses than observed in the developing female gonad. PDGF in the Sertoli cells is

negatively regulated by the gonadotropin FSH from the pituitary (Gnessi et al, 1995). Until 17.5

dpc, PDGFRα is only sporadically expressed throughout the testis and PDGFRβ is expressed

mainly in the interstitial tissue. By 18.5 dpc, gonocytes are also expressing PDGFRβ. There is

high expression of PDGF ligands and their receptors until PND 5 when the expression levels

drop drastically (Basciani et al, 2010). In the adult testis, the only significantly expressed PDGF

receptor/ligand combination is the alpha receptor and PDGF-AA isoform, expressed and

secreted primarily by the Leydig cells. Only in mice do Sertoli cells continue to produce PDGF in

adults (Mariani et al, 2002). PDGFRα and β are both found in the prenatal PMCs. In growing

and functional testis, PDGF-BB stimulates contraction of PMCs as well as cell proliferation.

PDGFR-BB also inhibits 5α-reductase and 3β-hydroxysteroid dehydrogenase activity in the

Leydig cells, stimulating increased testosterone production (Basciani et al, 2010). Postnatally,

PDGF is responsible for a number of actions in the testis such as increased testosterone

production, chemotaxis, proliferation and migration in PMCs, and germ cell proliferation and

differentiation. PDGF levels in the human testis are highest during the active foetal periods

between weeks 16 to 20 and significantly decreased during weeks 24 to 28. During the high

expression periods there is proliferation of gonocytes and foetal Leydig cells as well as

migration of the PMCs. At the end of the second trimester gonocytes begin their quiescent

period. PDGF levels again rise in the adult testis, evidence that it is also involved in

21

spermatogenesis. PDGF profiles were especially high in Leydig cell tumours (Basciani et al,

2002).

Testis development requires the communication and activation of different cell types

mediated through PDGF secretion from the Sertoli cells. Essentially Sertoli cells, through PDGF

production, are the conductors orchestrating the entire process of testis cord formation, and

without them, the testis would develop as an unorganized heterogeneous collection of cells.

Stimulation of organ cultures of 11.5 dpc urogenital ridge sections with PDGF-BB induces testis

cord formation through the MAPK and PI3K pathways. Alternatively, PDGF inhibitors block this

effect and repress the formation of seminiferous cords. Mesenchymal cells migrate from the

mesonephros to the gonad and develop into myoid cells, which play an integral role in testis

organization by surrounding the seminiferous cords and secrete contributions to the basal

lamina. Migratory mesonephric cells, myoid cell precursors, which also express p75

neurotrophin marker (p75NTP), are the only gonadal cells expressing PDGFRβ before 12.5 dpc.

PDGF-BB induces mesonephric cell proliferation, migration and chemotaxis through the

formation of cellular lamellipodia, a process that is vital for testis formation (Puglianiello et al,

2004). These cells in vivo migrate towards the testis around 13.5 dpc and enter the

interstitium. During this embryonic stage, PDGF-BB induces proliferation which causes

testicular growth and development. Migration of the testicular cells during development is a

“male specific event” which is regulated by the second messenger PI3K downstream to the

PDGF receptor (Ricci et al, 2004). PDGFRβ knockout mice die prenatally but generally develop

normally until 16-19 dpc. This drastic change appears to be due to defective PDGF-induced

migration of mesenchymal cells to their proper destinations. Without migration, vascular

smooth muscle cells and pericytes cannot develop appropriately (Puglianiello et al, 2004).

PDGFRα is also very important in testis development. Without PDGFRα, adult Leydig cell

production and function is impaired. Studies with PDGFRα KO mice showed that foetal Leydig

cells were practically normal while the adult Leydig cell population was almost non-existent.

This led to a significant decrease in testis size and spermatogenic arrest early in puberty.

Because foetal Leydig cells developed normally, embryonic testosterone levels were relatively

22

normal allowing masculinisation and testicular descent to proceed predictably. The PDGFRα KO

mice also caused a decreased expression in various other genes necessary for Leydig cell

development such as ptch1 (Brennan et al, 2003).

Imatinib mesylate is a tyrosine kinase receptor inhibitor that is used in cancer treatment

to block PDGFR and c-kit, a specific receptor for stem cell factor (SCF), attenuating the

proliferative activity of the tumour cells. C-kit, similar to PDGFR, is very important in testicular

development and is expressed on PMCs, Leydig cells, PGCs and migratory postnatal gonocytes,

while SCF is produced and secreted by the Sertoli cells. This factor is involved in the maturation

of spermatogonia and Leydig cells and migration of the gonocytes to the basement membrane.

While migrating gonocytes express c-kit, they lose the receptor after migration and only

express it again at the onset of spermatogenesis. In vivo treatment of PND 5 mice with imatinib

for three days caused a decrease in migration and proliferation of the gonocytes and decreased

induction of the stem cell pool, increased apoptosis in germ cells, and caused development of

shorter seminiferous tubules and lower testis weight. In adult mice that were treated

neonatally, there are increased levels of LH and FSH as a compensation for testis size and stem

cell count, but spermatozoa production is normal in the treated mice (Nurmio et al,

2007;Nurmio et al, 2008).

6.4 Effect of PDGF on Gonocytes

PDGF is not required for foetal gonocyte proliferation as germ cells do not express the

PDGF receptors before gestational day 13.5. PDGFRβ knockout studies show that in early

embryogenesis Knockout mice maintain a healthy germ cell population at this early stage of

development. It is hypothesized that Sertoli cells regulate the proliferatory and migratory stage

of the postnatal rat gonocytes, and later embryological foetal human gonocytes. It is well

known that PDGF is responsible for similar actions in many developing tissues and therefore it

was proposed that PDGF could be involved in this stage of testes development as well (Basciani

et al, 2008). Our lab has previously shown that rat gonocytes express PDGFRα and PDGFRβ

postnatally. At PND3 gonocytes also express a variant form of PDGFRβ in the cytosol which

contains no ligand binding domains but maintains tyrosine kinsase autophosphorylation

23

domains (Wang and Culty, 2007;Thuillier et al, 2003). Previous data from our lab proved that

PND3 gonocytes are stimulated to proliferate by both PDGF and 17β-estradiol, both of which

are produced and secreted by neonatal Sertoli cells (Li et al, 1997). These experiments were

conducted with pure gonocytes cultures in order to prevent any intercellular communication

with other testicular cell types. The proliferative effect of 17β-estradiol was inhibited by an

antagonist to estrogen receptors (Li et al, 1997). Foetal exposure to various estrogenic

compounds caused up-regulation of the gonocyte expression of PDGF receptor β (Thuillier et al,

2003;Wang and Culty, 2007), further suggesting an interaction between PDGF and estrogen

pathways in gonocytes. Treatment of mice between PND1 and 5 with imatinib confirmed the

involvement of PDGF-BB in gonocyte proliferation that our laboratory had previously described

in rats (Basciani et al, 2008). The inhibition of the PDGFR caused delayed gonocyte maturation

with a decrease in numbers as well. There was also an observed decrease in the migration of

gonocytes to the basal lamina of the tubules. Not only was there reduced proliferation but the

treated gonocytes also experienced an increase in apoptosis (Basciani et al, 2008).

6.5 V1-Variant Form of PDGFRβ

There are many known variant transcripts of the PDGFRs found in normal and cancerous

tissues, although their roles and functions remain puzzling (Mosselman et al, 1996;Mosselman

et al, 1994;Palumbo et al, 2002;Vu et al, 1989;Heinrich et al, 2003). We identified a variant

form of PDGFRβ in PND3 gonocytes named V1-PDGFRβ that is a cytosolic molecule missing part

of the extracellular ligand binding domain. It has active tyrosine kinase activity and is only

expressed at specific stages of testis development. V1-PDGFRβ transcripts are comprised of

intron 6 until exon 23. In the F9 teratocarcinoma cell line, V1-PDGFRβ was seen to play a role in

retinoic acid induced differentiation (Wang and Culty, 2007).

6.6 Pathologies Involving the PDGF Signalling Pathway

Increased PDGFR activity is linked to various pathologies, most importantly cancer.

PDGFs play a role in a number of different cancers including lung, prostate and renal. These

cancers can either use PDGF as an autocrine factor or through paracrine stimulation from other

secretory tissues. Most known gliomas express high levels of PDGF ligands as well as their

24

receptors. Another role for PDGF in cancer is stabilization of the vasculature through pericyte

recruitment, thus playing a role in angiogenesis of tumour blood vessels (Dai, 2010).

7. Extracellular-Stimulated Downstream Signalling Pathways

Extracellular stimulation of cells to elicit a critical change or specific cellular activity can

be initiated through two major pathways. Steroids can enter the cell and bind to their specific

receptor which will then activate one of hundreds of various cellular pathways. Most

extracellular signalling molecules or hormones cannot enter the cell due to their hydrophilic

properties and therefore must affect cellular functions from the outside. Substrate binding to

membrane bound receptors causes a conformational change in the receptor which elicits either

autophosphorylation of the intracellular domains of the receptor or stimulates specific GTPase

activity through G-protein coupled receptors. This enzyme activation recruits other proteins

which will then stimulate a cascade of secondary messengers to be phosphorylated or

dephosphorylated in order to direct a very precise response. Large scale amplification of

second messengers such as cyclic AMP or MAPK allow for strong responses to very minute

extracellular stimulations. Alternatively, scaffolding proteins function to maintain close

proximity between downstream enzymes or distinct communicating pathways to increase the

rate and specificity of the response. The nuclear receptor family response to steroid

stimulation will be discussed in depth in the following sections dealing with estrogenic cellular

activity. We will not be discussing G-protein coupled receptors, although they are equally

important in cell signalling. We will restrict our discussion of downstream pathways to the two

major pathways stimulated by growth factor tyrosine kinase receptors, and more specifically

the two pathways reported in PDGF induced testicular development (Ricci et al, 2004), the

mitogen activated protein kinase (MAPK) pathway and phosphatidylinositol 3-kinase (PI3K)

pathway.

7.1 Mitogen Activated Protein Kinase Pathway

The MAPK pathway is a linear downstream pathway cascade of kinase molecules

ranging from three to five molecules, each of which is phosphorylated by the previous member.

This cascade is involved in proliferation, gene transcription, migration, differentiation,

25

development, learning, survival as well as apoptosis (Robinson and Cobb, 1997). This is a very

highly conserved system of molecules and is present in the majority of eukaryotic cell types in

very high sequence conservation (Schaeffer and Weber, 1999;Kolch, 2000). The MAPK

molecule is the last enzyme in the cascade which carries out the necessary action, and as such

is phosphorylated by MAPK kinase, also called the MAPK/ERK kinase (MEK) which is in turn

phosphorylated by MAPKKK (or MEK kinase) such as Raf (Robinson and Cobb, 1997). MAPKs

are inactivated by the MAPK phosphatases (MKP) (Rumora and Grubisic, 2009). There are four

major known classes of MAPKs; extracellular signal regulated protein kinase (ERK), p38, c-jun N-

terminal kinase (JNK) (also known as stress activated protein kinase (SAPK) (Kim and Choi,

2010)) and BMK, each of which has a number of different isoforms or related pathways. The

ERK pathway is often involved in proliferation and is activated by growth factors (Rumora and

Grubisic, 2009) while JNK and p38 are commonly activated by stress factors such as tumour

necrosis factor α (TNFα), interleukin 1β (Il-1β) or cellular stress (Kim and Choi, 2010). It is

commonly observed that JNK and ERK have opposing actions in a single cell, one promoting

apoptosis and the other functioning in cell survival (Robinson and Cobb, 1997). The first

described MAPK, and the most researched pathway to date is ERK1/2 which is transcribed by

the erk1 and erk2 genes and are 42 and 44 kDa respectively (Shaul and Seger, 2007).

The first step of ERK activation is receptor ligand binding which triggers

autophosphorylation of the receptor on tyrosine residues. Ras, a small G-protein, is activated

and attaches to the receptor on SH2 (phospho-tyrosine) domains (Kim and Choi, 2010). Ras

activation is regulated by the RTK-Grb2-SOS complex coming to the membrane and docking to

the receptor. ERK can induce a negative feedback mechanism by phosphorylating SOS and

disassembling the whole complex (Kolch, 2000). This activation is also regulated by the

scaffolding proteins kinase suppressor of ras (KSR) or MEK partner 1 (MP1) (Kim and Choi,

2010). The specific enzymes that are activated for the phosphorylation of ERK1/2 in the

pathway are Raf and MEK1/2. Ras and Raf are very well known oncogenes, overexpression or

autonomous expression of which is responsible in part for a large number of cancers. Raf is a

cytosolic protein that is recruited to the phospholipid membrane by activated Ras where it is

activated through phosphorylation (Shaul and Seger, 2007;Kolch, 2000). For the process to

26

continue, MEK is required to bind to an active Raf molecule (Robinson and Cobb, 1997). MEKs

are very specific for Raf due to their proline rich domain, unlike all other MAPKKs. Upon

binding, MEKs are subsequently phosphorylated by Raf on two serine residues in the activation

loop (Shaul and Seger, 2007). Without this sequence, MEK cannot bind and therefore cannot

be activated by Raf (Schaeffer and Weber, 1999). MEK has many regulatory domains and

phosphorylation sites and can be controlled through a host of different enzymes including ERK

via a feedback mechanism which will either downregulate or upregulate MEK’s signal. MEK is

deactivated by the serine/threonine phosphatase PP2A. In the final tier of the MAPK cascade,

MEK phosphorylates ERK on a tyrosine and threonine residues in its activation domain. MEK

contains a docking site for ERK, enabling more precise and quick activation. Deactivation of ERK

is through dephosphorylation by PP2A, PTP-SL and MKPs. ERK is a serine / threonine kinase and

will phosphorylate substrates in the cytosol or the nucleus for activation. In the nucleus ERK

often activates the transcription factors Elk1m, c-fos, p53, Ets1/2 and c-jun (Shaul and Seger,

2007).

Both MEK and ERK are capable of translocation into the nucleus at rest, and at higher

frequencies when activated. Unlike ERK, MEK has a nuclear export signal (NES) and is exported

out of the nucleus via the exportin system upon entering. This is hypothesized to function in

activating nuclear-localized ERK molecules, or in recruiting ERK molecules that generally reside

in the nucleus. When inactive, MEK and ERK molecules are bound to docking proteins, ensuring

that they are, for the most part, found in the cytoplasm. When MEK or ERK is activated,

conformational changes occur, releasing them from their docking proteins, allowing for passive

or active (in the case of ERK, only as a homodimer) translocation into the nucleus. ERK is able

to dock itself to nuclear proteins when it is activated. In the nucleus, ERK is able to activate

transcription, proliferation and cell survival, while if activated ERK is recruited to the cell

membrane it can become a pro-apoptotic signal.

Specificity is a very important aspect of extracellular signalling. It is the objective of

rigorous research to understand exactly how this specificity and regulation occurs with such

precision. Additionally, it is important to understand exactly how it is possible that we get a

27

very specific response from the combination of various stimuli, activating pathways that often

can share similar molecules. Many molecules are ubiquitous among different pathways, but can

still be signal-specific in individual cells (Schaeffer and Weber, 1999). In PC12 cells, long-term

ERK phosphorylation led to cell differentiation, while shorter activation times led to

proliferation. Scaffolding proteins are very important as well to allow for proper cascade

activation. Scaffolds are integral in defining the cellular location of the ERK molecules and

forming multi-protein complexes, as well as coupling the MAPK signalling molecules with other

downstream pathways in a form of signalling crosstalk (Shaul and Seger, 2007).

The MAPK pathway is an active component of various pathologies. Its role in cancer, as

discussed above, is very important to our understanding of the disease and is the focus of a

considerable amount of ongoing therapeutic research. In addition, MAPK can function in

neurodegenerative disorders such as Alzheimer’s disease and Parkinson’s disease. In

Alzheimer’s disease, there is build up of amyloid-β plaque in the brain causing significant

memory loss and mental deterioration. One proposed mechanism is triggered by oxidative

stress forming reactive oxygen species (ROS) which activates JNK and p38. This triggers

neuronal destruction as well as stabilization of amyloid precursor peptide (APP) (Kim and Choi,

2010).

7.2 Phosphatidylinositol 3-Kinase

PI3K is a downstream pathway that also can induce proliferation, migration and

apoptosis. The PI3K pathway is based on the conversion of inositol lipids into signalling

molecules through phosphorylation. PI3K, which is made up of a p85 regulatory subunit and a

p110 catalytic subunit, phosphorylates phosphatidylinositol-4,5-bisphosphate (PI 4,5 P2 (PIP2))

to phosphatidylinositol-3,4,5-triphosphate (PtdIns (3,4,5) P3 (PIP3)) which then binds to either

phosphoinositide dependant protein kinase 1 (PDK1) or Akt via a PH (pleckstrin homology)

domain (domain that binds to triphosphorylated lipid inositols). This pathway can either be

activated by tyrosine kinase receptors or G-protein coupled receptors. There are other classes

of PI3K that activate different pathways as well. When Akt binds to the PtdIns (3,4,5) P3, it is

phosphorylated in its activation domain by PDK and by mTOR complex 2. Akt can now migrate

28

to the rest of the cytosol or the nucleus. There are over 100 substrates that Akt can affect in

the cell (Martelli et al, 2010). In certain cell types, PI3K actually functions as an activator for the

MAPK pathway. In mesangial cell proliferation and migration, PI3K induces MAPK response, and

ERK is turned off when the cells are inhibited by a PI3K inhibitor (Choudhury et al, 1997). As

part of my thesis research, our lab has recently showed that PDGF-induced PND3 gonocyte

proliferation is only through the MAPK pathway and not via the PI3K downstream cascade

(Thuillier et al, 2010). (See Results and discussion sections)

8. Estrogen

8.1 Endogenous Estrogens

Estrogens (oestrogens) are classically known as the “female sex hormones”, as their

function in the female reproductive system well preceded knowledge of their function in other

systems. Estrogens are crucial for the development, maintenance and function of the female

secondary sex characteristics. Found in their highest concentrations in females of reproductive

age, they function in developing breasts and the endometrium, and play a role in regulating the

menstrual cycle. In addition to their sexual functions, estrogens regulate growth, bone

mineralization, brain masculinisation and cardiovascular functions in males as well as females

(Luconi et al, 2002); and they also play a role in the immune system and the central nervous

system (Heldring et al, 2007). In the absence of estrogen, bones cannot mineralize and the

epiphyseal plate cannot close. For this reason, estrogen deficiencies often lead to very tall

stature and continued growth with a high incidence of osteoporosis. Estrogens play a

protective role in the cardiovascular system by inhibiting vascular smooth muscle cells from

proliferating, thereby preventing vascular injury response as well as preventing the

development of atherosclerosis, a disorder due to lipid deposits in the large arteries (Luconi et

al, 2002). Three main types of endogenous estrogens are found in the endocrine system; 17β

estradiol (E2), estrone and estriol, E2 being the strongest and most potent of the three.

Estrogens are steroid hormones, and therefore are derived from cholesterol, a molecule

combining four five-carbon and six-carbon rings. They are secreted as endocrine messengers

and therefore circulate through the bloodstream until reaching their target tissue where they

29

will bind the estrogen receptors (ERs), triggering the expression of specific genes regulated by

estrogen (Heldring et al, 2007). The enzyme aromatase, generally found in the endoplasmic

reticulum of estrogen secreting tissues, catalyzes the irreversible conversion of androgens such

as testosterone to estrogens. Aromatase is an enzymatic complex containing a ubiquitous

reductase and a cytochrome P450 aromatase which harbours a heme and a steroid binding

pocket (Carreau et al, 2003).

What defines a molecule as an estrogen? Originally scientists used to test possible

estrogen candidate molecules by implanting them into the uterus of female rats or rabbits to

see if there was a change in cell proliferation. If growth occurred, this suggested that the

molecule was estrogenic, therefore estrogens were considered uterotropic molecules (Koehler

et al, 2005). Today there are two in vitro methods performed at the molecular level using

either recombinant ERs or gene reporter assays, which are more precise than the latter in

defining estrogenic molecules. These new techniques have provided scientists with a new

means of defining estrogenic molecules. The first method is a competitive binding assay testing

the new compounds for their ability to displace the binding of E2, the most common estrogen,

from ERs. There are two main types of ERs, ERα and ERβ that will be discussed below in greater

detail, and both types of recombinant receptors are used for testing the possible estrogenic

activity of compounds. One can use fluorescently conjugated molecules to view binding of the

unknown chemical compared to E2 to find out how strong their relative affinity is to the

receptors. If they do bind, they are thought to have estrogen-like properties. The second and

more definitive method uses a luciferase-reporter-gene assay. The substance to be tested is

incubated with cells from a cell line that contains ER and an estrogen-dependent luciferase

gene containing an ER binding site on its promoter sequence. If this gene is transcribed in the

presence of the chemical, then the molecule can be considered estrogenic. If the chemical

binds to the ER, but does not induce gene transcription, it is regarded as an antiestrogen

(Gutendorf and Westendorf, 2001).

30

8.2 Exogenous Estrogens

There are many different molecules called xenoestrogens, created by humans or

originating from other natural sources such as plants that are present in the environment that

can bind and affect the activity of ERs, (Nikov et al, 2001). They are usually ingested or

absorbed into the body and are often found in our water sources. Although they act like

estrogens, their chemical structure doesn’t always appear similar (Nadal et al, 2000). These

chemicals are found in nature in a very diverse array of structures and can bind to the ERs with

different affinities, displaying estrogen-like functions (Nikov et al, 2001). Natural or synthetic

exogenous compounds that interact with the normal ERs are called endocrine disrupters and

display either estrogenic or antiestrogenic properties, either functioning similarly to natural

estrogens or inhibiting the transcriptional function of ERs (antiestrogens) (Gutendorf and

Westendorf, 2001). Some xenoestrogens are produced naturally by plants such as soy

(phytoestrogens), while others are synthesized as industrial chemicals, for example; pesticides

and herbicides, which become environmental pollutants. Generally, these molecules will have

a common structural motif containing a phenol attached to a bulky hydrophobic structure. Due

to different structural properties, some of these substances will have higher affinities to the

ligand binding domain (LBD) of the receptors than others. AdDP1 appears to bind more strongly

to both ERs than AdP2, and AdMP3 does not bind to either protein. These molecules have

reduced binding affinity compared with DES4 and 4OHT5

1 4,4’-(1,3-adamantanediyl)diphenol

, although AdDP will bind to ERβ with

higher affinity than 4OHT (Nikov et al, 2001). DES is one of the most potent xenoestrogens that

was used to prevent miscarriages in the 1940’s to 1970’s, until it was associated with birth

defects in males and females alike. Another extremely well-known xenoestrogen is bisphenol-A

(BPA) which is usually found in canned food, plastic bottles and dental sealants (Nadal et al,

2000). A well known ER antagonist is ICI 182780 which is often used in research to examine

whether or not ER binding is required in the specific effect of an estrogenic compound. It

2 Adamantly substituted phenol (4-(1-anamantyl)phenol) 3 2-(1-adamantyl)-4-methylphenol 4 diethylstilbestrol 5 4-hydroxytamoxifen

31

competitively blocks the receptors, and allows scientist to see if the estrogenic effects continue

in the absence of a functioning ER (Delbes et al, 2006). ICI 182780 is also called Faslodex or

Fulvestrant and is used clinically to treat cancer (AstraZeneca, 2004). Another family of

compounds interacting with ER are the pharmacological modulators called Selective ER

Modulators (SERM), such as raloxifen and tamoxifen that have differential affinity for the two

ERs. SERMs can be agonists for one receptor and antagonists for the other (Luconi et al, 2002).

Many of these molecules disrupt the endocrine signalling of the receptors, and depending on

the dosages and period of exposure can either cause or prevent different diseases including

cancer. Alternatively, other exogenous estrogen compounds might also be able to prevent

cancer in the right doses. One of the xenoestrogens believed to function this way is the most

common phytoestrogen in soy, genistein, which has a very high affinity to ERβ and therefore

slows down proliferation in some tissues. Other synthetic estrogens are used clinically for the

same purpose (Heldring et al, 2007).

8.3 Estrogen Receptors

8.3.1 Genomic Function of the Estrogen Receptor

In order for the Estrogens to perform their biological functions, they must travel to the

cells in the targeted tissues and subsequently bind to their specific receptors. There are two

known estrogen receptors (ERs), each transcribed from different chromosomes (10 and 12 in

mouse, 1 and 6 in rat and 6 and 14 in humans (Delbes et al, 2006)), namely ERα and ERβ (Segars

and Driggers, 2002). Estrogen receptors were first discovered in the 1950’s by Elwood V.

Jensen who identified the ERα protein (Jensen, 2004). ERβ was discovered much later in 1996

by Jan-Åke Gustafsson while looking for an androgen receptor in prostate cells (Maher, 2006).

ERs function as transcription factors (Segars and Driggers, 2002) and are part of the steroid

nuclear receptor superfamily, which functions in regulating gene expression (Delbes et al,

2006). After ER binds directly to its specific ligand, it releases associated receptor inactivating

heat shock proteins (usually HSP90). Binding to estrogen causes the ER to change its

conformation allowing it to form a stable dimer with another activated receptor of the same

type. The dimer then translocates through the nuclear membrane into the nucleus (Luconi et

32

al, 2002). ERs can bind directly to DNA on selected estrogen response elements (ERE) found

near or in the promoter region of the targeted gene. In order to fully activate the transcription

of their target proteins, the ERs must also come in contact with coregulatory proteins that can

either enhance or repress their functions (Segars and Driggers, 2002). It is the ER dimer that

interacts with the ERE and attracts other molecules to the promoter to either suppress or

activate gene transcription (Nikov et al, 2001). ERs can also be modified by phosphorylation

(Segars and Driggers, 2002). Because transcription is regulated by hormones, the pattern of the

genes that are modulated depends on all of the signalling pathways that are active at the time

of ER activation in the targeted gene (Heldring et al, 2007).

8.3.2 Estrogen Receptor Structure

Each of the two ERs has several isoforms, which are truncated versions of the wild-type

protein, some due to differential splicing of the C-terminal. There are no reported functions of

the ERα variant isoforms, although there is a lot of research being conducted concerning

different functions of truncated versions of ERβ (Delbes et al, 2006). Human ERα is a protein

encoded by 9 exons, and is built up of 595 amino acids and a molecular weight of 66kDa. ERβ is

smaller at 530 amino acids and 54kDa, but is also encoded by 9 exons. Both receptors have the

structure of a common nuclear receptor (NR) with six functional domains. The N-terminal is the

least conserved domain (Luconi et al, 2002). The central domain or C domain contains two zinc-

fingers creating the DNA binding domain (DBD). The zinc finger motifs are coordinated by eight

cysteine residues (Pettersson and Gustafsson, 2001). This domain is the most evolutionarily

conserved and is very important for the binding of the receptor to the DNA (Heldring et al,

2007). The DBD is connected to the ligand binding domain (LBD) through the D domain, which

acts as a hinge between the two domains. The hinge domain is not very conserved between

different NRs and associates itself with HSP90. The LBD, or E/F domain, is multifunctional as it

binds to ER agonists or antagonists. It also plays a role in dimerization, transactivation, cofactor

binding, and can bind to a second HSP90 molecule (Pettersson and Gustafsson, 2001).

Transcription is activated in the receptor through two activation functions (AF-1 and AF-2) on

either end of the receptor molecule, both of which accepts and binds to coregulatory proteins

33

(Heldring et al, 2007). While both ERs share similar affinities for E2, they have varying affinities

for other natural and synthetic estrogen molecules as well as other various antagonistic ligands

(Luconi et al, 2002). There are two common isoforms of ERβ; ERβ1 and ERβ2. ERβ2 has 54 extra

nucleotides, 18 of which are inserted into the LBD (Pettersson and Gustafsson, 2001). It

appears that ERβ2 is a suppressor of ERβ1 and ERα (Luconi et al, 2002) and interestingly it is

found in human foetal gonocytes (Gaskell et al, 2004).

8.3.3 Non-Genomic Effects of ERs

Recent research has shown that in addition to the transcriptional function of ERs,

estrogens are capable of triggering a very rapid cascade of responses. These effects are much

faster than those produced by ER working on its own accord and can happen within seconds to

minutes following stimulation. These responses are activated through second messengers such

as calcium, activated kinases and tyrosine kinases, PKA6 and PKC7

8.4 Reproductive Effects of Estrogens

as well as the ERK pathway.

These effects, unlike those generally activated by estrogens through the classical ER

mechanism, occur in the cytoplasm or on the cellular membrane. These pathways are possibly

triggered by different functions of the known ERs or perhaps a completely new type of ER

(Luconi et al, 2002). An estrogen-activating membrane bound G protein-coupled receptor has

been found in breast cancer as well as on adult testis (Delbes et al, 2006). Nongenomic effects

of estrogens will be discussed in more depth with more recent research and clinically relevant

examples in the following sections on signalling crosstalk mechanisms.

8.4.1 Estrogen Signalling in Females

In the female reproductive system E2 is synthesized in the ovaries by granulosa cells via

aromatization of testosterone. Estrogens are also made in adipose tissue, skeletal muscles,

skin, hair and bone. Women of reproductive age produce the highest concentrations of

estrogens; post-menopausal women are more susceptible to heart disease and bone fracture,

6 Protein Kinase A 7 Protein Kinase C

34

due to their low levels of estrogen production. Hormone replacement therapy was thought to

be very reliable in decreasing these risks, but has recently been shown to pose other health

risks such as increased risk of stroke in postmenopausal women and is now recommended only

in specific cases for treatment of menopause (Anderson et al, 2004). In adult females, ERβ is

found in the granulosa cells where its levels of expression fluctuate throughout the menstrual

cycle. It is also present in the glandular epithelial cells. ERα is abundant in the theca cells as

well as luminal, glandular epithelial cells and stroma. Women deficient in estrogens develop

normally until puberty, but then show failure in growth of the breasts, enlargement of clitoris

and unfused epiphyses as well as amenorrhea. Such patients, when treated with estrogens and

progesterone begin normal puberty cycles (Pettersson and Gustafsson, 2001). Estrogens play a

crucial role in the early development of the female reproductive organs as they directly

stimulate secondary sex characteristics, breast development, uterine development and

maturation of the fallopian tubes (Waterloo, 2007).

8.4.2 Estrogen Signalling in Male Reproductive System

ERs were first located in the male reproductive system over thirty years ago in the

epididymis, almost twenty years after Jensen’s discovery of the estrogen receptor. It was then

apparent that estrogen played a role in male foetal organ development, but not much else was

known. Even as late as the 1990’s some scientists continued to think that the ERs present in

the male reproductive system were simply the residue of embryological differentiation (Hess

and Carnes, 2004). Due to differences in laboratory methods of research, there are differing

opinions as to which receptors and isoforms are found in which tissues and cells. It appears

that ERβ is more prevalent in the accessory organs, and is expressed mainly in the prostate,

bladder, seminal vesicles and testis (Luconi et al, 2002), while ERα is mainly in the efferent

ductules and Leydig cells (Hess and Carnes, 2004). Both receptors are found in the epididymis

and spermatogonia as well as elongated spermatids (Luconi et al, 2002). Our own studies

identified ERβ in germ cells. Including late foetal to neonatal gonocytes, and in older germ cell

stages (Wang et al, 2004). There are many ligands other than the common endogenous

estrogens that are able to bind to ERs, as they are, unlike many other nuclear receptors, not

35

highly specific receptors, and may accept a large number of different molecules with a

surprising diversity in structure (Heldring et al, 2007).

8.5 Estrogen in Male Reproductive Development

ERs are found in the foetal testis from a very early stage of development. ERα is found

in the mouse gonad 10.5 days postconception and is prominent in Leydig cells only until birth.

ERβ is seen in the foetus a few days after ERα and is generally located in the gonocytes, Sertoli

and Leydig cells. With the knowledge of estrogen production in the foetal and neonatal testis

in addition to the presence of ERs, it was anticipated that estrogens played an important role in

foetal development. It is now known that estrogen is important in the regulation of the adult

male reproductive system as well. By inactivating specific estrogen receptors, scientists

showed how estrogens functioned in the developing testis. In the foetus, excess estrogens

inhibit the proper development of the testis. ERβ is involved in gametogenesis since mice

expressing a mutant inactive ERβ presented increased gonocyte numbers from late foetal to

neonatal periods, due in part to decreased germ cell apoptosis. However, estrogens appear to

exert either transient proliferative effects in vivo (Vigueras-Villasenor et al, 2006) and on

isolated gonocytes (Li et al, 1997;Thuillier et al, 2010), no effect on gonocytes used in organ

cultures or co-cultures, and negative effects on foetal gonocyte numbers (Delbes et al, 2007).

However, ERβ knockout male mice are fertile (Gould et al, 2007;Krege et al, 1998) suggesting

that ERβ is not essential for male germ cell development or that redundancy occurs in the germ

cells of male transgenic mice due to the presence of another yet uncharacterized estrogen-

dependent receptor. In this context, PGCs were shown to express the estrogen related

receptor beta (ERR-β) which was found to play a role in their proliferation (Mitsunaga et al,

2004). Studies in our lab have shown that estrogen combined with PDGF enhances gonocyte

proliferation, most probably via ERβ which is strongly expressed in gonocytes (Li et al,

1997;Thuillier et al, 2010). It has also been shown that ERα inhibits foetal germ cell growth,

indicating a delicate balance between the two functions in fetal development (Carreau et al,

2006). Another possible function of estrogens in the developing testis might be to establish the

adhesion between Sertoli cells and germ cells (Hess and Carnes, 2004;MacCalman et al, 1997).

36

ERα was proposed to mediate the negative effects of estrogens on foetal Leydig cells (Delbes et

al, 2004;Delbes et al, 2006). After puberty, estrogens have been shown to play a role in the

regulation of spermatogenesis, spermatid maturation, germ cell number and viability (Carreau

et al, 2003). By injecting small doses of estrogens into animals, scientists were able to

accelerate the onset of spermatogenesis.

In rats, plasma estrogen levels rise throughout gestation and peak at 17.5 days, while

the mother’s plasma estrogen levels are highest at 18 days gestation. Also located in the foetal

testis is the estrogen producing enzyme aromatase indicating that estrogens can be

manufactured in the testicular tissue. Aromatase is most prominent in foetal Sertoli cells. Due

to this, they are probably the cells producing the most estrogens in the foetal testis. However,

Leydig cells were demonstrated to be the primary estrogen producer cells in adult testis (Delbes

et al, 2006). The major aromatase molecule found in the testis is cytochrome P-450 aromatase

(CYP19) (Wahlgren et al, 2008). Testicular estrogen levels are generally much higher than

plasma levels, indicating that the majority of estrogens produced in males are synthesized by

the testicular tissue (Hess and Carnes, 2004). Aromatase has been found in germ cells, which

appears to be another large producer of estrogens. Aromatase also plays a role in elongated

spermatid mobility (Nitta et al, 1993;Berensztein et al, 2006;Carreau et al, 2003). A study

conducted by Wahlgren et al (2008), demonstrated that estrogen can induce DNA synthesis and

a proliferatory effect not only in developing male germ cells, but in adult spermatogonia as

well. 5α Androstane-3β, 17β-diol (3βAdiol) is a specific ERβ agonist derived from 5α-

dihydrotestosterone (DHT) secreted by Leydig cells, and is found in elevated levels in the rat

testis. In vitro studies show that 3βAdiol was able to induce DNA production in premitotic

spermatogonia, evidence that estrogens are important in spermatogenesis (Wahlgren et al,

2008). Estrogens are crucial for the development and maintenance of the male reproductive

system and must be in perfect balance with testosterone in order for normal development to

take place. Too much or too little estrogen can be harmful to development; therefore it is clear

that “endogenous estrogens are essential for maintenance of male fertility” (Delbes et al,

2006).

37

8.6 The Estrogen Hypothesis

Increasing concern has been generated in the medical community throughout the past

fifty years regarding research documenting a gradual increase in severe male reproductive

disorders. As society is becoming more industrialized, it is hard not to notice the detrimental

effects that it has on the well-being of many different organisms in the environment. Studies

have shown that over this period of time there has been an unprecedented upsurge in

testicular cancer, cryptorchidism and hypospadia (Delbes et al, 2004) as well as a large

reduction in average male sperm counts from 170 to 70 million spermatozoa per millilitre

(Delbes et al, 2006). Both cryptorchidism, a malfunction in testicular descent, and hypospadia,

the abnormal positioning of the urethra on the penis, are associated with increased likelihood

of cancer. The cause of these findings was first hypothesized by Sharpe and Skakkebaek in 1993

to be environmental, when they stated that the current situation is due to an escalating

quantity of xenobiotics and xenoestrogens in our environment, making the risk of exposure

very likely. They stated that “estrogen-like molecules could alter adult male fertility by acting

early in gonad development and that inappropriate exposure to estrogen during foetal or

neonatal life could affect adult reproductive function”. Irrefutable evidence for this statement

can be observed in the reproductive development of males exposed to DES in their foetal life.

This group of male offspring, most of who were born between the 1940’s and 1970’s, had an

abnormally high incidence of reproductive disorders in their adult life (Delbes et al, 2004).

Many of these changes to the endocrine and reproductive system have been shown to be

completely irreversible. Exposure to exogenous estrogens appears to affect females as

negatively as it does in males (Iguchi et al, 2001).

There are many other occurrences in nature that strengthen this hypothesis and have

prompted focused research in this field. Alligators in a chemically polluted lake in Florida

showed a very large upsurge in abnormalities to their reproductive organs during development

(Gutendorf and Westendorf, 2001) compared with those alligators in a nearby lake unpolluted

by agricultural waste containing xenoestrogens. The affected alligators had low testosterone

levels and micropenis with abnormally developed testicular tissue (Delbes et al, 2006). Fish in a

38

lake near a sewage disposal site also displayed a decrease in sperm quality. Trout in Lake

Ontario died due to the dioxin-like compounds in the water (Gutendorf and Westendorf, 2001).

The most extreme example observed in human society is apparent in the Aamjiwnaang First

Nation community in Canada, where over the course of ten years the gender ratio of viable

offspring saw drastic changes whereby only 30% of the children born to this community were

male. This phenomenon was suggested to be due to the nearby industrial and environmental

chemicals (Delbes et al, 2006).

In light of the given data and the numerous examples of estrogen-induced disorders, it is

still very unclear how these estrogenic compounds induce their negative effects on the male

reproductive system, and whether or not there actually is a direct estrogenic effect at all.

Certainly, it is well documented that estrogen plays an important role in foetal and neonatal

testicular development and an imbalance of estrogens is a cause for concern, and similarly it is

important to address the fact that unrealistically high dosages of in utero estrogenic exposure

can cause serious mal-developed reproductive organs, however the majority of studies

experimenting with normal, potential daily exposure levels of exogenous estrogens have only

observed minimal, benign changes in the reproductive system. Nevertheless, there is a concern

surrounding potentially harmful estrogenic compounds found in everyday life such as the

phytoestrogen genistein in soy products and BPA, which can be found in plastic containers.

Testicular estrogen exposure is even further complicated by the fact that it has been shown

that at different stages, estrogens can be pro-proliferative or pro-apoptotic and each of these

functions have been linked directly to the estrogen receptor pathway. In order to further

understand the mechanisms behind changes caused by foetal estrogen exposure, our lab

conducted various experiments, observing estrogen-induced changes in the PDGFR, estrogen

receptor and MAPK pathways in foetal, neonatal, prepubescent and adult rats.

Estrogen receptor β in gonocytes are complexed together by a host of chaperone

proteins including Hsp90, the cochaperone p23 and the immunophilin cyclophilin 40 (Cyp40),

which stabilize the receptor in the cytosol, protecting it from degradation and exposing it for

ligand binding. When estrogen binds to its receptor, Hsp90 is dissociated but p23 remains

39

attached as ERβ translocates to the nucleus. In utero exposure to BPA, genistein, coumestrol

and DES increase mRNA and protein levels, at 21dpc and PND3, of Hsp90, but not those of p23.

DES is the only estrogenic compound that increases Cyp40 levels in the neonatal testis. Unlike

the other estrogens that increase Hsp90 in the gonocytes, BPA increased Hsp90 levels in the

whole testis. These changes were only found in neonatal rat testis, as Hsp90 levels were

normal in PND21 prepubescent rats. The only notable change at PND21 was a significant

increase in spermatogonia (Wang et al, 2004), although by PND60 in adult rats, spermatogonial

numbers were back to normal (Thuillier et al, 2009). Exogenous estrogen exposure to foetal

rats induced an increase in PDGFRα and β mRNA as well as protein in PND3 rats compared to

controls. There was a notable increase in PDGFRβ in gonocytes. DES exposure gave the most

intriguing result, as it increased PDGFRβ in the seminiferous cords at lower doses, but then

actually decreased it when exposed to a higher dosage. Foetal rats showed a significant

increase in PDGFβ in exposed testis (Thuillier et al, 2003). These results are very important,

leading to our understanding that estrogen-induced proliferation is somehow modulated

through the PDGF pathway (Thuillier et al, 2010). In addition to this, our lab has recently

observed that rats exposed to BPA and genistein in utero caused increased levels of ERK1/2 and

Raf1 in PND3 rat testis. Similar to previous findings, these changes were no longer present in

PND60 rats, although it was noted that PND60 rats had an increased number of Leydig cells,

although testosterone levels in the testis were found to be normal. At experimental levels of

exogenous estrogens similar to potential environmental exposures, there was no significant

decrease in fertility observed in treated adult rats (Thuillier et al, 2009).

8.7 Approaches Used to Study the Role of Estrogen/ERs in Males

8.7.1 Laboratory Results of Estrogen Exposure in Vivo

In view of the natural models mentioned above, scientists set out to understand how

exogenous estrogens affected organisms in vivo. In order to gain an understanding of the

mechanisms used by estrogens in the male reproductive system, scientists studied animals at

different stages of development following exposure to large amounts of estrogens as compared

to unexposed animals. It appears that most of the testicular disorders originate from exposure

40

during foetal development. For instance, an irradiation-induced decrease in the number of

gonocytes during development was shown to lower sperm count later in life (Moreno et al,

2001). When viewing data on patients exposed to DES as a foetus, it was interesting to find

that some patients were unharmed, while others had highly deformed reproductive organs. It

was understood that the differences in intensities of reproductive deficiency were based on the

stage of pregnancy in which they were exposed. DES appears to have a greater negative effect

during the first semester of pregnancy (Delbes et al, 2006). Some of the malformations

observed while exposing foetal mice to exogenous estrogens were low testicular weight, low

germ cell count, increased apoptosis in germ cells, and a high rate of tumours (Delbes et al,

2004). In vitro studies showed that DES exposure in early development disrupted cord

formation as well as establishment of many of the reproductive cells. Pregnant mice were

given exogenous estrogens either by gavage, injection or ingestion via their drinking water in

order to expose foetal mice. The results of these studies displayed both short term and long

term repercussions in the affected mice (Delbes et al, 2006). Some of the short term effects

included accelerated testicular development and abnormal gonocyte differentiation (Delbes et

al, 2004). It is conclusive though, that foetal development is very sensitive to exogenous

estrogens (Delbes et al, 2006). Although it is now clearer that estrogens can play a role in

testicular development, this is not proof that the small amounts of estrogens that we are

exposed to chronically actually contribute to testicular diseases. Whereas most experimental

effects of estrogen are detrimental to the reproductive organs, there have also been studies

describing positive effects of specific estrogens as shown in germ cell development. Indeed,

one study showed that exposure to mild doses of genistein actually increased sperm’s fertility

(Delbes et al, 2004). Some of the natural or synthetic non-proliferative or antiestrogen

molecules are now used to treat breast cancer as they prevent the cells from continuously

multiplying. The first attempt to use anti-estrogens to prevent cancer was through the use of

an anti-ERα molecule. However over time the tumour grew resistance to this molecule. For

this reason scientists have attempted to create an ERβ agonist to treat cancer instead, in view

of the pro-differentiation and/or anti-proliferation role of ERβ in prostate and breast tissue

(Heldring et al, 2007). Indeed, ERβ ligands were shown to slow prostate tumour growth in nude

41

mice and promise to lead to the development of novel and efficient therapeutic approaches

(Koehler et al, 2005).

8.7.2 Knockout Mice

One method of understanding how a specific gene or protein works in the body is to

study the animal in the complete absence of its expression and/or function. By inactivating a

target gene, we can see the functioning differences between the wild-type animal and its

mutant. The most common technique used in laboratories is the knockout mouse; a mouse

that is created with the specific gene encoding the target protein deleted from its genome.

Many studies have been performed using ER knockout (ERKO) mice to better understand the

role of estrogen in males and females. The first completely ER inactive ERKO mouse was

produced in 1996 by Korach et al. in order to better understand ER’s function in the male

reproductive system. This breakthrough preceded the discovery of ERβ, and therefore in

actuality was only an ERαKO mouse. At this point in time it was assumed that the whole ER

function was removed from these mice, however with the discovery of ERβ, it was soon

understood that only ERα was affected, leaving ERβ fully functional in the original ERKO mouse.

Korach and his team made their ERKO mouse by disrupting the ERα gene by the use of

homologous recombination. The heterozygous mice were mated in order to obtain progeny

homozygous for the deletion of the ER gene. In order to understand the function of ER, they

performed long and short term mating assays with wild-type, heterozygous and ERαKO mice

and then euthanized them to see the anatomical differences between the different clones. In a

two month mating period the wild-type and heterozygous male mice produced equal numbers

of viable offspring, while the ERαKO mice did not produce a single litter. It was also noted that

in the short term mating (one day per week for three weeks), the ERαKO mice produced almost

no copulatory plugs, while the males in the other two categories seamed to mount the females

at almost every opportunity. After euthanasia, testis weight was much lower in ERαKO mice,

while the other organs did not differ much from the wild-type. ERαKO mice also had higher

testosterone levels. The epididymides of ERαKO mice contained low numbers of sperm after 20

weeks of life and the testis contained atrophic and degenerating seminiferous tubules. At birth

42

the ERKO testis appeard quite normal but as the mice grew older, more degeneration occurred

in their tubules, and their lumens became increasingly dilated. Furthermore, ERαKO mice

presented a much higher percentage of immotile sperm, which again increased with age. In

vitro fertilization showed that even the motile sperm were less fertile than wild-type

spermatids. This study clearly showed that ERαKO mice were completely infertile and the lack

of ERα affected sperm number and function, as well as a suggested behavioural change in the

mouse’s sex drive. Estrogen appears to be involved in regulating luminal fluid production in the

Sertoli cells, which is why ERKO mice produced an overabundance of the fluid and the

seminiferous tubules were dilated. Another cause might be due to inhibited reabsorption in

the efferent ducts of the testis due to loss of estrogen function causing fluid accumulation.

Fluid accumulation might be an important explanation as to why the seminiferous tubules

degenerate over time. As fluid accumulates in the testis, intratesticular pressure increases,

blocking off regular blood flow to the surrounding tissues. As the pressure rises, the less

vascularised tissues degenerate first, followed by the sections with more vasculature (Eddy et

al, 1996). There is also an increase in cryptorchidism in ERαKO mice, similar to the observations

made following high levels of exogenous estrogen exposure (Luconi et al, 2002).

Although ERα is important for testicular function it is not responsible for

spermatogenesis. By implanting ERαKO germs cells into an empty wild-type testis, normal

sperm is produced and the cells can generate live offspring. ERβ probably does influence

spermatogenesis as diets high in soy increase sperm production in adults and genistein has a

much higher affinity to ERβ than ERα. However, studies in rats exposed neonatally to genistein

showed that small numbers of these rats were sterile in adulthood compared to none in control

animals, suggesting that neonatal exposure can have deleterious effects in a subset of animals

(Atanassova et al, 2000). Additionally, long term exposures to ICI 182,780 cause gradual

testicular atrophy. This effect was thought to have to do with fluid accumulation, but that was

only the case in rats and not in mice. Thus, estrogens may have a direct effect on the ERβ

molecules on germ cells, where ICI 182,780 functions as an antiestrogen (Hess and Carnes,

2004).

43

ERβKO mice were created similarly to ERαKO mice by inserting a neomycin resistance

gene into the third exon of the ERβ gene using homologous recombination. The results were

quite different than those of the ERαKO mice. Unlike the female ERαKO mice that were

completely infertile, these females were fertile and breast development was normal. The only

effect on females was that their litter size and number was smaller than for the wild-type. Male

ERβKO mice were fertile as well, and their testicular tissue showed no abnormalities, although

older males had epithelial hyperplasia in the prostatic ducts and bladder wall (Krege et al,

1998). A later study showed that homozygous and heterozygous mice carrying an inactivating

mutation to ERβ expressed an increase in gonocyte number and density, although the actual

testis size remained the same. There was also a decrease in gonocyte apoptosis. From this

study, it was concluded that endogenous estrogens normally directly inhibit germ cell growth

during foetal development. Estrogens are crucial for testis development and germ cell

proliferation, although ERβ does not affect Leydig and Sertoli cells (Delbes et al, 2004). It has

now also been shown in the male prostate that there is an abundance of 3βAdiol, the second

most prominent estrogen in the body, which is a very good ligand for ERβ, allowing it to prevent

proliferation in the prostate. ERβKO mice also showed that many cells in the prostate do not

differentiate completely and are therefore still capable of proliferation (Koehler et al, 2005).

A third strain of knockout mice was created to further understand the function of

estrogen in males. These mice were deficient in the gene that produces the aromatase

enzyme, rendering them unable to produce estrogens. Male mice of this breed developed

normally and reproduced until they reached five months of age, at which point they

progressively lost their spermatogenesis function to become deficient of round spermatids and

completely infertile by one year of age (Carreau et al, 2003). At this point, the testis also

became deficient of round spermatids (Hess and Carnes, 2004). The reason for the inhibition of

spermatogenesis in ArKO mice is different than in ERαKO mice, and is due to a complication in

germ cell differentiation. Estrogen is used as a survival factor for normal spermatids, and when

this hormone is not present, they cannot differentiate and therefore begin to apoptose.

44

Experiments on mice defective for both ER molecules (ERαβKO mice), showed results

similar to those of the ERαKO mouse. Although from these results it appears that the only truly

crucial receptor is ERα, it is important to note that we are still not perfectly clear on the

function of ERβ and therefore it is hard to see what it does based on knockout models. It also

might be the case that, although we have knocked out the target receptor or protein that we

believe is important, there may be alternative molecules that perform the functions of the ERs

or of the aromatase enzyme in their absence (Luconi et al, 2002). Female ERαβKO mice

displayed a distinct phenotypical difference from those only deficient of ERα. Whereas it is

clear that in females there is a specific function for ERβ, it has yet to be completely defined in

males (Couse et al, 1999). From the results using the different knockout mice, it appears that

the two estrogen receptors can often play opposite roles in various experimental situations. It

may be crucial for the body to have the proper balance between the two of them in order to

develop and function correctly (Heldring et al, 2007). Although a distinct functional role for ERβ

has yet to be defined in male testicular development, and more specifically in the

differentiation of the male germ cells and reproductive organs, it is clear that due to their high

expression in testicular cells at this developmental period there must be a crucial

developmental role for ERβ.

9. Cell Signalling Cross Talk Mechanisms

Studies in our lab have demonstrated that unlike PND2 gonocytes, PND3 gonocytes

respond to PDGF as well as 17β-estradiol by proliferating in a non-additive manner, suggesting

the existence of crosstalk between the two signalling pathways. PDGF and 17β-estradiol are

likely activators for the regulation of the second proliferative phase in vivo given that both

factors are found at high levels in rats until PND5 or 6. Testosterone, the male reproductive

hormone, which is synthesized by Leydig cells in early neonatal life, could also play a role in

germ cell development and differentiation, although these cells are generally considered to be

androgen receptor deficient (Johnston et al, 2001). Because our initial study suggested that

PDGF and 17β-estradiol might crosstalk to stimulate gonocyte proliferation, we decided to

determine which downstream cascade of PDGF was involved in this process by using inhibitors

45

and antagonists specific for either PDGF-activated signalling molecules or ER. My master’s

thesis work is part of this study, which will be presented in details in the results and discussion

sections.

There are a number of models in the literature of different downstream pathways

working together in order to activate a single cellular function. As mentioned above, there are

many novel estrogen-induced functions, above and beyond transcription of genes expressing

the ERE in their promoter region. Roles in proliferation and rapid effects of estrogens have

stimulated a new field of research in secondary and non-genomic mechanisms of estrogen/ER

binding. Similarly, as it has become clear that one of the most important regulatory

mechanisms of the MAPK pathway is its interaction with other downstream pathways, it is

important to study where these interactions take place and how to manage them in

pathological situations. It is becoming increasingly evident that many known cellular actions

actually require some degree of physical or chemical crosstalk between different pathways,

including nuclear receptor pathways. The models in place are valuable tools in our

investigation of PND3 gonocyte proliferation and, using our experimental data, we can begin to

hypothesize how this crosstalk works and at which point in the respective downstream

pathways it takes place.

9.1 Intercommunication of Separate Downstream Pathways

Analogous to computer programming, where if the right buttons are not pressed and

the settings are not in complete coordination the program will not run, cell signalling requires

all of the right inputs and stimuli to be activated simultaneously or in the correct sequence of

events to elicit the proper response. Increasing research has shown that each cellular function

is actually the result of many factors all working in communication. In specific cells, high levels

of cyclic AMP (cAMP) can turn on or turn off Raf molecules, to direct a specific response upon

tyrosine kinase receptor stimulation. There is a small GTPase designated as Rap1 that responds

directly to cAMP that can activate or inhibit specific Rafs in a dose dependant, cell specific

manner (Schaeffer and Weber, 1999). In this way, the same growth factor will have a different

response based on the levels of cAMP in the cell at the time of stimulation. There are small

46

molecules such as small G-proteins and others that can communicate between different MAPK

pathways to increase or attenuate a response to an extracellular stimulation. There are

transcription factors and other important cellular enzymes that are affected either positively or

negatively by more than one MAPK. This allows for different stimulations to have additive

affects on the specific factors (Schaeffer and Weber, 1999). Raf or MEK have been shown to be

activated by cyclins or cAMP which then stimulates a normal MAPK cascade. P38, a MAPK, can

activate phospholipase A2 in platelets upon stimulation by collagen (Robinson and Cobb, 1997).

9.2 Crosstalk between Estrogen Receptor and Growth Factor Receptors / MAPK

Pathway

Nongenomic estrogen activity generally works alongside numerous different pathways.

In contrast to the classical localization and function of estrogen receptors, ERs have been

detected on cell membranes, denoted as membrane estrogen receptors (mERs), and activate

downsteam pathways. There are numerous different hypotheses as to what these receptors

are and how they work. It is assumed via immunofluorescent cytology that there are some

systems and cell types expressing the ordinary ERα and ERβ docked to the membranes, while

there have also been novel ERs discovered as well, most notably, the ER-X receptor found in the

brain, uterus and lungs (Hirahara et al, 2009) and GPR30 which is an estrogen binding G-protein

coupled receptor (Fox et al, 2009). Very often, these systems express different splice variants

of ER that help with the cellular response to estrogen (Hirahara et al, 2009). mER behaviour

resembles that of G-protein coupled receptors, activating G-protein GTPase activity (Simoncini

et al, 2006), and mERs are usually found interacting with growth factor receptors or activating

MAPK cascades on their own. Interaction with growth factor receptors can be either on the

same cell as the mER or on a neighbouring cell in a paracrine or juxtacrine crosstalk (Osborne et

al, 2005). Through this mechanism, ERs can activate pathways including MAPK, PI3K, GSK-3β,

Src, STATs nitric oxide synthesis and intercellular calcium flux (Collins and Webb, 1999;Hirahara

et al, 2009). Src is an important molecule, as ER has no kinase activity; most estrogen induced

phosphorylation of other pathways is modulated by this kinase (Fox et al, 2009). ERα has an

Sh2 domain and can bind directly to Scr to activate MAPK (Geffroy et al, 2005).

47

There are a number of proposed mechanisms for how ERs can dock to the plasma

membrane given that they have no transmembrane domains in their amino acid sequence. ERs

have been shown to be recruited to the membrane and dock to growth factor receptors via

activation of docking proteins. Alternatively, ERs can attach to the end of the cytoskeleton near

the plasma membrane using caveolin-1 as a docking protein. This complex can directly target

MAPK for activation (Hirahara et al, 2009). Caveolin has also been shown to be upregulated

upon estrogen receptor activation (Park et al, 2009). Wong et al. published a paper in 2002

describing a protein called modulator of non-genomic activity of estrogen receptor (MNAR)

which complexes with ERs and Src as a scaffolding protein to regulate estrogen-induced

proliferation. Although the author retracted this paper in 2009, the scaffolding complex is still a

valid mechanism of mER localization and activation (Wong et al, 2002), and other groups have

reported similar results in their studies (Osborne et al, 2005). It was reported by Cheskis et al.

that MNAR is integral in ER’s stimulation of the MAPK pathway via phosphorylation of MNAR by

c-Src. c-Src is activated by dephosphorylation of its inhibitory phosphotyrosine sites or by

binding to receptors via its Sh2 and Sh3 domains. It is not yet known whether or not ER

requires a scaffolding protein to activate PI3K (Cheskis et al, 2008) although it is known that ERs

somehow bind to PI3K to activate Akt (also called protein kinase B, PKB) (Geffroy et al, 2005).

Other scaffolding proteins are also reported to form complexes with ER to enable the non-

genomic estrogen responses are Shc, IGF-1R (Fox et al, 2009) and MTA1-S (Osborne et al, 2005).

Fu et al. showed in bovine arterial endothelial cells that estrogen-induced proliferation

was caused by expression of the cell cycle modulator cyclin D1 through the ERK pathway. This

ER was found to be bound to the membrane and is likely to be a novel ER. The promoter region

of cyclin D1 contains binding sites for NF-κB, AP-1 and Sp1, all of which can be activated by ERK.

This study has important implications for cardiovascular (CV) disease, as postmenopausal

women tend to be at a higher risk for suffering from CV disease due to estrogen deficiency. The

finding that estrogen induces endothelial cell proliferation is an important revelation in

understanding this phenomenon and possible remedies (Fu et al, 2007). Oligodendrocytes and

myelin cells also express mERs, and estrogen induces proliferation of these cells as well.

Estrogen deficiency is a common risk factor for neuron demyelination and multiple sclerosis,

48

and treatment with estrogen has been shown to produce some positive results (Hirahara et al,

2009). Osteoporosis is another disease that is frequent in estrogen deficient postmenopausal

women. Bone, which is in constant remodelling, requires estrogenic activity for proper strength

and structure. An imbalance of estrogen can result in destruction of calcified bone. Although it

has been proposed that estrogen hormone therapy could reverse osteoporotic changes,

estradiol therapy is not recommended (Sehmisch et al, 2008). Instead, due to the increased risk

of breast cancer with estrogen supplementation, it was proposed to use phytoestrogens

instead. The phytoestrogen resveratrol, which comes from such plants as nuts, berries and

grapes, appears to be a promising treatment for osteoporosis by stimulating estrogen-induced

proliferation and maturation of osteoblasts through the ERK and p38 pathways (Dai et al, 2007).

Resveratrol has been shown to prevent development of osteoclasts through RANK-L, and

induces apoptosis in these cells (He et al, 2010). There are groups working on resveratrol

analogues with more potent effects on osteoclast and osteoblast differentiation (Kupisiewicz et

al, 2010), as well, the phytohormone 8-prenylnaringenin has been shown to have a much

stronger anti-osteoporotic effect than resveratrol (Sehmisch et al, 2008).

Bouskine et al. demonstrates how estrogens can stimulate proliferation in the JKT-1

testicular germ cell tumour cell line through ERK and protein kinase A (PKA) activation. ERK and

PKA then activate the transcription factor cAMP response element binding protein (CREB). It

was proposed that this does not go through the classical ERs as its function is inhibited by G-

protein inhibitors as opposed to estrogen antagonists. Although this study concluded that

inappropriate foetal exposure to estrogen can cause germ cell hyperplasia leading to testicular

cancer, it was also noted that in these cells ERβ acts as a tumour suppressor gene and prevents

cancer progression (Bouskine et al, 2008). Crosstalk-induced proliferation in cancer cells has

commonly been observed between estrogen receptors and growth factors. Normal mammary

growth requires communication between estrogen and epidermal growth factor (EGF) while

both estrogen and IGF-1 are necessary for uterine tissue proliferation. When any of these

factors are overexpressed, this leads to hyperplasia of the tissue. Proliferation in many breast

cancer models is stimulated by a rapid estrogenic response through MAPK. Proliferation is

often through the STAT5 molecule which is activated by Src or EGF receptor (Fox et al, 2009). In

49

endometrial carcinoma cells, treatment with estrogen induces phosphorylation of ERK, and an

influx of calcium into the cell. When estrogen binds to its membrane receptor, mER stimulates

the opening of L-type calcium channels generating an influx of calcium into the cell. A

conceivable mechanism for this action is ER stimulation of phospholypase C through a G-

protein. In this model, ERK activation is indirectly associated to estrogen receptor activation

through changes in the calcium electrochemical gradient (Zhang et al, 2009).

Growth factors can also be used in eliciting an estrogenic response. In endometrial

adenocarcinoma cells, EGF activates nuclear ER-induced transcription of genes through specific

EREs. EGF ligand binding caused a downstream pathway that activates ER in the nucleus to

induce transcription (Ignar-Trowbridge et al, 1995;Ignar-Trowbridge et al, 1992). Ginsenosides,

a potent phytoestrogen derived from ginseng root is proposed as another possible treatment

for postmenopausal effects such as cardiovascular and central nervous system diseases.

Unfortunately, ginsenoside also induces breast cancer growth and development. This

phytoestrogen stimulates a normal genomic response as well as a non-genomic estrogenic

response. It induces proliferation of MCF7 breast cancer cells via stimulation of a tyrosine

kinase receptor through the MAPK pathway which activates ERK. ERK is required to

phosphorylate ERα on serine 118 in the nucleus in order to fully activate it (Lau et al, 2009).

Activation of ERK is involved in a very intricate crosstalk mechanism whereby reactive

oxygen species, secreted in very low concentrations by spermatozoa, self-stimulate themselves

activating a number of different signalling pathways which are all required to activate and

upregulate protein tyrosine phosphorylation for spermatozoa capacitation prior to fertilization

of the oocyte. The main crosstalk involves the ERK signalling cascade in parallel with the PKA

downstream pathway through phosphorylation of downstream molecules. The mechanism

proposed is that ERK activation works in parallel with adenylyl cyclase, cAMP, PKA and its

downstream molecules, and they converge together for the phosphorylation of protein

tyrosine. This crosstalk is set in place to ensure the coordinated precise timing of spermatozoa

capacitation directly before fertilization (Awda and Buhr, 2010;de and O'Flaherty,

2008;O'Flaherty et al, 2006).

50

Estrogens have been shown to induce increased PDGF and PDGFR expression in uterus

and vagina, demonstrating that estrogens can affect PDGF-dependent responses by increasing

PDGFRs and their ligands in specific tissues (Gray et al, 1995). An observation made in the

MCF7 breast cancer cell line is that progesterone, through its specific receptor, induces PDGF-

AA production and secretion. Although the PDGFR in many breast cancer cells is down

regulated, the PDGF-AA was shown to play a paracrine role by increasing proliferation and

viability of vascular smooth muscle cells in vitro in a crosstalk through PDGFRα. Progesterone

stimulation of the cancers cells induces communication with the endothelial cells to promote

vascularisation for the tumour to grow. This is important, given that tumour progression

requires the continued growth of blood vessels through vascular smooth muscle cells.

Targeting the progesterone pathway could provide a successful treatment for these types of

cancers (Soares et al, 2007). Testosterone and DHT are other critical steroid hormones in male

reproductive tissues that have been shown to crosstalk with growth factors or other steroid

hormone pathways. For example, AR was found to interact with ERs via Src in a human

prostate carcinoma cell line (Migliaccio et al, 2000) and it was shown to crosstalk with a number

of growth factor pathways such as that of IGF-1 (Kaarbo et al, 2007;Culig et al, 1994). Because

testosterone is produced by the foetal Leydig cells still present in neonatal testis in rodents, one

could consider the possibility that it might also act on gonocyte functions (Habert et al, 2001).

Although most germ cell stages do not express androgen receptor (AR) and are therefore not

considered as direct androgen targets (Vornberger et al, 1994), a recent study of the Habert lab

suggested that AR might be transiently expressed in foetal gonocytes where androgens were

found to inhibit proliferation (Merlet et al, 2007). This led to experiments I performed in which

the effects of testosterone and progesterone on gonocyte proliferation were examined (see

results and discussion sections, Thuillier et al. 2010).

10. Summary

Previous findings in our lab have shown that PDGF-BB and 17β-estradiol are able to

stimulate PND3 gonocyte proliferation, through activation of their respective receptors. This

was confirmed by in vitro experiments where AG370, a specific inhibitor of the kinase activity of

51

PDGFR, attenuated the PDGF-BB-stimulated proliferation, and ICI 182780, an ER antagonist,

arrested the estrogen-induced proliferative effect. The results obtained with ICI 182780

confirmed those observed earlier with a different ER antagonist, ICI 164384, showing that the

17β-estradiol effect requires binding on an estrogen receptor (Li et al, 1997). These earlier

experiments also showed that adding AG370 together with 17β-estradiol on gonocytes blocked

the proliferative effect on estrogen, while adding ICI 182780 with PDGF-BB prevented PDGF to

stimulate gonocyte proliferation.

In view of the knowledge listed above and the earlier findings in our laboratory, the goal

of my research was to extend and clarify our understanding of the mechanisms involved in the

stimulation of gonocyte proliferation by PDGF and 17β-estradiol, focusing on identifying the

downstream molecules mediating this process. My hypothesis was that PDGF and estrogen are

working in concert via crosstalk to activate gonocyte proliferation, and are both in some way

involved in ERK1/2 activation to transform the cell from the quiescent state to the proliferative

phase. My goal was to identify which signalling pathway was involved in the PDGF effect and

whether it was also under the control of estrogen.

52

11. Materials and Methods

11.1 Gonocyte Isolation

Male albino Sprague Dawley rats were purchased at post natal day 2 (PND2) from

Charles Rivers Laboratories. The pups were euthanized by inducing hypothermia and

decapitation at PND3 according to the guide for the care and use of experimental animals from

the Canadian Council on Animal Care and McGill University. Neonatal rats are resistant to CO2

asphyxia and were therefore euthanized by decapitation. Dissection was performed by

abdominal incision to remove and isolate the undescended testes from 30 pups. Testes were

kept in a 50mL tube of RPMI 1640 supplemented with 100 U/ml penicillin, and 100 mg/ml

streptomycin, on ice for the duration of the dissection. The testes were manually decapsulated,

and then processed for tissue dissociation. The cell isolation procedure was executed according

to the protocol described by Li et al. (1997), which is a modified version of that used by Van

Dissel-Emiliani et al. (1989). Under sterile conditions, the decapsulated testes were incubated

with 3.8mL RPMI, 2.5mL type IV Collagenase (stock 2mg/mL), 1mL Hyaluronidase (stock

1mg/mL) and 0.8 mL DNase I (stock 1mg/mL) in a shaking water bath at 37℃ for 30min in order

to cause the seminiferous tubules to dissociate from the interstitium. After decantation of the

digest, the pellet of remaining tissues containing the seminiferous cords was processed for

further tissue digestion, while the supernatant, including a mixture of mainly dissociated Leydig

cells, vascular endothelial and blood cells was discarded. The tubule fragments were

dissociated by adding 3 ml of 0.25% trypsin + 1 mM EDTA and 0.1 ml of DNAse I for 15 min at

37℃ to dissociate the Sertoli cells, peritubular myoid cells (PMCs) and gonocytes. The

digestion was arrested by adding 5 ml of 10% foetal bovine serum (FBS) (which neutralizes

trypsin) in RPMI. After collection of the supernatant, the remaining undigested tissue was

incubated a second time with trypsin-EDTA and DNase I, and stopped by addition of 10% FBS.

The final cell suspensions were filtered through a 40 μm nylon filter to remove any cell

aggregates or tissue fragments. The gonocytes were quantified with a hemacytometer,

differentiated from the other cell types by their larger size and round, smooth morphology.

The cells were collected by centrifugation at 800g, the pellet was resuspended in RPMI and the

53

X 30

Hya

luro

nida

se

Colla

gena

seSe

min

ifero

usTu

bule

s

Deca

psul

atio

n

Tryp

sin, E

DTA

Tubu

les a

nd se

para

ted

cells

Filtr

atio

n

Ove

rnig

ht in

cuba

tion

Gono

cyte

s, S

erto

li an

d M

yoid

cells

BSA

grad

ient

m

ixer

4%2%

BSA

Gra

dien

t

Frac

tion

Colle

ctio

n

Purif

ied

Gono

cyte

s

2:15

hr

Sedi

men

tatio

n

54

Figure 4: Gonocyte Isolation Protocol. Schematic representation of thedissection and digestion of rat testis for the purpose of isolating gonocytes, theprespermatogonial germ cell pool in neonates. Testis are collected in ice-coldRPMI and then manually decapsulated to allow the tissue to be exposed to thedissociation enzymes. Collagenase and hyaluronidase are used to dissociatethe interstitial tissue from the seminiferous cords, while maintaining the cordsintact. This digestion removes the Leydig cell population from the sample, aswell as the vasculature and other interstitial cell types. The next digestion iswith trypsin-EDTA, which breaks down the cords and dissociates the cells fromtheir tight intercellular adhesions. We now have a cell suspension of somaticcells and 0.5-1% gonocytes. Overnight incubation allows Sertoli cells and myoidcells to adhere to the culture dish, removing 50% of the somatic cells from thesuspension. Sedimentation by gravity on a BSA gradient allows the cells tofraction out by size. Fractions of 2mL are collected and observed on thehemacytometer to determine the numbers and purity of the gonocytes in eachfraction. Fractions with the highest purity of gonocytes are pooled togetherand used for our experiments. Other sets of gonocytes at lower purities aresometimes also kept and used for different investigations.

55

cells were cultured overnight in 150 mm large culture plates with 5% FBS in RPMI in 5% CO2 at

37℃. This incubation was crucial to gonocyte isolation as Sertoli and myoid cells are adherent

to culture dishes but gonocytes are not. This allowed us to separate out 50% of the

Sertoli/myoid cells before applying the cells to a BSA gradient. The cells were collected after 20

hours in culture, centrifuged and resuspended in 2mL RPMI, and then fractionated through a 2-

4% BSA gradient. The cells were allowed to sediment by gravity for 2 hours and fifteen

minutes, approximately 30x 2mL fractions were collected, and the number of gonocytes

determined on a hematocytometer. The fractions with the highest number and purity of

gonocytes compared with somatic cells were pooled together and centrifuged. Fractions of

lower purity that still contained a high yield of gonocytes were pooled separately. The total

yield of the high purity fraction was between 2 X 105 and 4 X 105 gonocytes at 45 to 70% purity.

The lower purity fraction often yielded similar numbers of gonocytes at purities ranging from

20-40%. Different purities were selected depending on the quantity of cells yielded from a

single dissection as well as the type of experiment being performed, with western blotting

requiring higher percentages of gonocytes and cytochemistry allowing for lower purities. The

final cell suspensions were resuspended in RPMI without FBS and incubated with or without

treatments for 30 and 60 minutes at 37℃.

11.2 Cell Culture for Short Term Molecular Profiles

For cytological studies, between 8,000–15,000 cells per well were treated for short term

assays of 30 and 60 minute cultures in RPMI, without FBS. FBS contains many hormones and

growth factors in it and therefore for short term assays, it was important to have no external

factors acting upon the cells other than the treatment itself. Cells were also cultured with FBS

for 4 and 8 hours to observe translational and early transcriptional changes in protein

expression as well as 20 - 24 hours for proliferative and gene transcription studies.

Alternatively, for some proliferative assays, cells were also cultured with charcoal stripped

(Biomeda) or PDGF-depleted serum. When conducting cultures for western blot analysis,

25,000–30,000 cells were treated in each well or tube. Because these studies were short term

analyses, most of the experiments were conducted in microcentrofuge tubes and the cells were

56

incubated in a 37℃ shaking water bath as opposed to using cell culture plates. Treatments

included 10ng/mL human PDGF-BB, 10-6M 17β-estradiol, 100µM AG370 (Biomol Research

Laboratories), 100µM ICI 182780 (Tocris) and 10µM UO126 (Calbiochem), between 10-6M and

10-15M of Genistein (synthetic) and Bisphenol A (4,4’-isopropylidenediphenol; BPA;

ICN/Millipore Biomedicals). Cells were also treated with 10-6M testosterone and 10-6M

progesterone. Although cells were in sterile tubes, the water bath was not sterile, thus cells

were monitored for viability and contamination. Cell viability was excellent, and no cultures

showed any sign of contamination.

11.3 PDGF-depleted FBS

FBS was prepared overnight at 4℃ with specific antibodies against PDGF-AA, BB and AB

(Biodesign International and Oncogene research products). In order to remove antibody-PDGF

complexes, the FBS was run through a protein A HiTrap column that binds antibodies together

with their antigens. Removal of most of the PDGF was documented with Western blot analysis.

11.4 Protein Analysis – Western Blot

Western blot analysis was carried out in order to a) determine gonocyte expression of

cell signalling molecules, growth factors and hormone receptors as well as germ cell markers,

and b) investigate changes in expression of these factors when gonocytes were treated with

PDGF and estrogen in the short term. Expression of these molecules provides clues for the

mechanisms behind gonocyte proliferation as well as the commitment of gonocytes to

differentiation. Changes in expression can illustrate which factors are activated or expressed

after short term exposure to PDGF and estrogen, leading to an understanding of the crosstalk

between the two distinct pathways.

The cell cultures were stopped after 30-60 minutes by addition of 10µM Protease

inhibitor cocktail (Sigma P8340), 10µM phosphatase inhibitor (Roche, PhosSTOP – 1 tablet for

10mL) and 4µM okadaic acid (Calbiochem, 495609) and centrifugation at 4℃ for 2 minutes at

5000g. The supernatant was removed and the cells were lysed and stored in a Laemmli buffer

containing 1X DTT including protease and phosphatase inhibitors as well as bromophenol blue

57

as a loading dye. Proteins were quantified using a BCA assay and spectrophotometer. Protein

samples were loaded onto a 4-20% or 10% tris-glycine (Invitrogen) gel and proteins were

separated via SDS-page electrophoresis. 20-30μg of protein or protein from 20,000-30,000

cells was applied to each well in the electrophoresis gels. Following electrophoresis, the

samples were transferred onto a nitrocellulose membrane using 1X Novex Tris-Glycine Transfer

Buffer (Invitrogen LC3675, 25X stock) in 20% methanol. Membranes were blocked with 5%

powdered milk or 5% BSA and incubated overnight with specific primary antibodies at

concentrations of 1:250 to 1:1000 in 1X TTBS at 4℃. The primary antibodies included

antibodies specific for ERK and phosphorylated-ERK (p-ERK) (PhosphoPlus MAPK kit from Cell

Signalling #9100), Raf1, PDGFRβ, MEK1, PI3K (from Santa Cruz #SC227, SC432, SC219, SC7189),

MIS/AMH (R&D MAB1146), ERβ (ABR PA1-313), ERα (Upstate #06-935), and PDGFRα (Upstate

#07-276). Secondary antibodies were incubated at 1:10,000 dilutions and were coupled with

streptavidin-horseradish-peroxidase (HRP) (Zymed Invitrogen Corp.). Loading controls for these

samples were histone-1 and Tubulin (Chemicon MAB052 and Abcam ab0742-300 HRP bound).

Negative controls were performed by incubation with rabbit and mouse IgG instead of primary

antibody or by using a secondary antibody alone, prior to reprobing the membranes with

specific primary antibodies. ERβ antibody signal was effectively blocked by incubation with the

specific peptide used to produce the antibody. The immunoreactive bands were visualized

using ECL-Plus detection reagents (Amersham). The images were collected and densitometry

analysis performed on a FujiFilm Luminescent Image Analyzer (LAS-4000) and Multi Guage V3.0

software. Using normalization against tubulin and histone gave quialitativly similar results but

higher variability due to the inconstancy in the Histone and Tubulin immunoreactions between

different experiments. Within each experiment, the same amount of total protein per sample

was loaded onto the polyacrylamide gels. In some of the experiments we verified, by using two

different protein concentrations on the same gel, that the immunoreactions of the proteins of

interest were proportional to the amount of protein loaded on the gel. Therefore we chose to

directly compare the immunoreactive bands of treated samples with the control and expressed

the data as a fold change for each protein analyzed. This allowed us to better express the

average changes observed between numerous experiments.

58

11.5 Nuclear Isolation

Nuclear extractions were performed to provide insight into which of the downstream

signalling molecules were activated inside the nucleus or translocated into the nucleus at the

time of activation following PDGF and estrogen treatments.

The cells were incubated for 30 or 60 minutes at 37℃ with the different reagents, then

protease and phosphatase inhibitors were added and the cells were centrifuged for 2 minutes

at 5000g and the supernatants were removed. The samples were kept on ice for the duration

of the cell fractionation and the cell pellets were resuspended with tissue homogenizing buffer

(with added protease and phosphatase inhibitors) and a cell fractionation buffer (0.25M

sucrose, 25mM KCl, 50mM Triethanolamine, 5mM MgCl2, 0.5mM PMSF, 1mM DTT). Cells were

transferred to a manual glass cell homogenizer tube and were lysed with a Teflon coated tissue

grinder. The cell lysates were centrifuged at 800g in order to pellet the nuclei, while cytosolic

proteins and small organelles (mitochondria, ER…) remained in the supernatant. The nuclear

fraction was resuspended in Laemmli buffer. The supernatant was transferred to a 10kDa

microcon filter (Millipore, YM-10) to concentrate the proteins in the supernatant. Most of the

remaining buffer was removed by centrifugation at 4℃ for 1 hour at 14,000g, and the

remaining protein concentrates were collected in a microcentrofuge tube and further

resuspended in Laemmli buffer.

11.6 Immunocytochemistry

Immunocytochemistry was used to localize the various signalling molecules expressed in

the gonocytes that could be activated during proliferation. These experiments helped localize

important signalling players and possible crosstalk mechanisms that could be involved in

proliferation.

Isolated gonocytes were kept for 30 or 60 minutes in microcentrofuge tubes in a 37℃

water bath. The reaction was stopped with 4% paraformaldehyde combined with phosphatase

inhibitors, protease inhibitors and okadaic acid for 7 minutes. The cells were then centrifuged

for 3 minutes at 5000g and resuspended in PBS. Cytological slides of the gonocytes were

59

prepared by cytospin for 10 minutes at 300RPM. The slides were fixed in a 60:40% acetone

methanol mixture for 7 minutes and stored at −20℃. Immunocytochemistry was performed

by exposing slides to 10% Dako-Cytomation Target Retrieval solution (Dako North America Inc.)

at 95℃, then blocking with PBS with 10% BSA and 10% goat serum. Primary antibodies were

applied overnight at 4℃ in 1:25 or 1:100 dilutions. The secondary antibodies were

fluorescently-coupled rabbit and mouse Alexa 488 and 546 antibodies from Invitrogen, and

were incubated with the slides for 1 hour in the dark at 1:300 dilutions. A nuclear (4’,6’-

Diamidino-2-phenylindole, DAPI, Invitrogen) fluorescent stain was applied and the slides were

mounted with Permafluor aqueous mounting medium (Thermo Scientific) and covered with a

glass coverslip. The slides were observed via fluorescent or confocal microscopy. Negative

controls were performed using mouse and rabbit IgG as well as secondary antibody alone.

Immunocytochemistry was used to investigate protein expression of Raf1, MEK1, ERK1/2, p-

ERK1/2, MIS/AMH, PI3K, ERα, ERβ, PDGFRα, and PDGFRβ. Studies were conducted to observe

the expression of certain proteins in control gonocytes as well as to investigate the short term

effects of protein expression and localization following the given treatments.

Photomicrographs were taken of the treated slides focusing the objective to an optimal DAPI

staining to ensure that the picture is taken in the middle of the cell, and intensity of the protein

expression was quantified using ImagePro 6.3 analysis software.

11.7 Proliferation Assay

Cells were cultured with one or more of the following reagents: PDGF-BB, 17β-estradiol,

xenoestrogens, testosterone and progesterone, with or without the specific inhibitors AG370,

UO126 and ICI 182780. The cells were incubated overnight together with 30µg/mL 5’-Bromo-

2’-deoxyuridine (BrdU) and 3µg/mL 5-fluoro-2’-deoxyuridine as described in the cell

proliferation kit from Amersham. Following cell culture and cytospin, the slides were fixed and

stained with a biotinylated anti-BrdU antibody (Neomarkers or Exalpha). The cells were

counted and positive staining was calculated as a percent of the total number of cells. PCNA

was used in some experiments to compare to BrdU in determining the percentage of

proliferating cells. PCNA gave similar results to BrdU.

60

11.8 Immunohistochemistry

Important for understanding the whole picture of PND3 gonocytes, expression of

important factors and receptors were investigating using 5 μm testes paraffin tissue sections.

Paraffin sections of PND3 testes were stained for protein expression of Raf1, MEK1,

ERK1/2, PI3K and MIS/AMH. Two methods were used for deparaffination and antigen retrieval.

Trilogy solution was used in an electric pressure cooker according to the company’s

instructions. Alternatively, slides were rehydrated in xylene and different dilutions of alcohol

and were then heated to 95℃ with DAKO antigen retrieval solution and blocked with 1% BSA,

10% goat serum and 0.02% Triton X100 to prevent nonspecific antibody binding. Primary

antibodies were diluted in PBS with 1% BSA, 0.02% Triton X100 and were incubated at a 1:100

dilution overnight at 4℃. Fluorescent secondary antibodies were used and the nucleus was

stained with either Hoechst 33342 (Invitrogen) or DAPI. The sections were mounted with

ProLong Antifade solution and were viewed under a fluorescent microscope. The same

negative controls were applied on the tissue sections as for the cytospin slides.

11.9 V1-PDGFRβ Vector Transfection and Live Cell Imaging

Green fluorescent protein (GFP)-tagged V1-PDGFRβ variant was transfected into

gonocytes to a) see if we could transiently overexpress a GFP-labelled protein in gonocytes and

b) to monitor V1-PDGFRβ expression and localization in treated cells to observe if V1 plays a

role in short term induction of gonocyte proliferation. Rats were dissected at PND2 and

incubated overnight with a V1-PDGFRβ constructed plasmid. The plasmid was constructed by a

previous student (Wang and Culty (2007) who previously characterized V1-PDGFRβ. The coding

sequence included the complete coding region between exon7 and exon23 of the PDGFRβ

gene. The plasmid was prepared in opti-MEM with a 1:1 ratio in Lipofectamine. Sertoli cells

from the dissection were also plated and incubated on glass confocal plates to prepare a Sertoli

cell matrix for the gonocytes. Following incubation, gonocytes were further purified on a BSA

gradient. They were plated on the Sertoli cell matrices and incubated overnight again in RPMI,

without serum, with V1-PDGFRβ plasmid. This allowed the gonocytes to settle into the matrix

of Sertoli cells and adhere to the somatic cells, also providing more time for the plasmid to be

61

transfected into the cells. After overnight incubation, gonocytes were treated with PDGF-BB

and 17β-estradiol in different combinations with and without inhibitors, and were monitored

for 60 minutes using live cell imaging by confocal microscopy. Short term profiles of the V1-

PDGFRβ expression was observed and photomicrographs were taken.

11.10 Statistical Analysis

The statistical analysis for this study was performed using Microsoft Excel’s

mathematical functions unpaired, two-tailed t-test. Each of the treatments was compared

directly to the control subject, as well as treatments were paired up against each other for

statistical analysis as well. The histograms presented in this study were assembled using

GraphPad Prism (version 5).

62

12. Results

12.1 Charcoal Stripped FBS and PDGF-Depleted Serum

Our first step was to verify that normal FBS was able to provide a steady basal level of

both PDGF and estrogen that would explain why each individual agent was able to stimulate

gonocyte proliferation without addition of the other one. For this, gonocytes were incubated

overnight with PDGF or 17β-estradiol in the presence of either charcoal stripped FBS (void of all

steroids), or PDGF-depleted FBS (figure 5). Gonocytes in culture with charcoal stripped FBS did

not display PDGF-induced proliferation, while cells treated in PDGF-depleted serum also could

not induce estrogen-stimulated proliferation. When 10-6M 17β-estradiol was added to the

charcoal stripped serum, treatment with PDGF was again able to induce proliferation. Similarly,

when PDGF was added to the PDGF depleted serum, estrogen-induced proliferation was re-

established.

12.2 Gonocyte Expression of Downstream Molecules of the PDGF Signalling

Pathway

Studies performed earlier in our lab identified the mRNA expression of Erk1, Raf1, Mek1

and PI3K through in situ hybridization in PND3 rat gonocytes. Erk1 had the highest mRNA

expression while PI3K mRNA was in much lower quantity. To confirm that gonocytes express

the corresponding proteins of these transcripts, we investigated their protein expression using

immunohistochemistry of PND3 normal rat testes. Raf1, Mek1, ERK1/2 and PI3K were all

expressed in the seminiferous cords, specifically located in the cytoplasm of Sertoli cells and

gonocytes, and once again PI3K signal was faint (figure 6A). The Sertoli cell marker AMH/MIS

was also present in the seminiferous cords of the testis sections. We then performed

immunocytochemical analysis to confirm these findings in isolated gonocytes using cells

collected and immunostained immediately after gonocyte isolation from PND3 rats. These

experiments confirm that ERK1/2, Raf1 and Mek1 were expressed in gonocytes (figure 6B). It

was also noticeable that of the morphologically smaller somatic cells present with the

gonocytes, some stained positively for all these factors, while others did not. There were also

AMH positive and AMH negative cells, which most probably distinguished between Sertoli cells

63

*

** **

FBSCharcoal-striped FBS

PDGF17β-estradiol

+---

-+--

+-+-

-++-

+-++

-+++

50

40

30

20

10

0Pr

olife

ratio

n ra

te (%

)

***

***

FBSPDGF-depleted FBS

PDGF17β-Estradiol

*** ***

+---

-+--

+--+

-+-+

+-++

-+++

50

40

30

20

10

0

Prol

ifera

tion

rate

(%)

**

Figure 5: PDGF- and estrogen-induced proliferation in the presence of serum devoid ofsteroids or PDGF. Gonocytes isolated at PND3 were treated with PDGF and estrogen ina medium containing either normal serum, PDGF- or steroid-depleted serum. The cellswere incubated overnight with BrdU, and the proliferating BrdU-positive cells werecounted to calculate their percentage against the total gonocyte numbers. Each valuerepresents proliferation rates ± SEM and was calculated from 3 experiments. For all ofthe histograms in this paper, statistical analysis was performed using a two tailed T-testand the significance values are given as follows: *: P<0.05, **: P<0.01, ***: P<0.001.

64

and PMCs. It is interesting to note that there were some gonocytes expressing AMH, most

likely due to AMH uptake by gonocytes following Sertoli cell secretion. Gonocytes were

examined after one day in culture following isolation, and we were able to confirm the

continued expression of these downstream molecules. The strongest signals were from ERK1/2

and Mek1. Finally, Western blot analysis also confirmed the presence of these downstream

proteins, and we showed that the more prominent ERK isoform in PND3 gonocytes is ERK1

which is expressed ten times higher than ERK2 (figure 6C). Somatic cells had a higher

expression of the proteins examined in total protein extracts as compared to gonocytes (figure

6C).

Given the results above, our lab set out to determine which of the two downstream

pathways was involved in PDGF-induced proliferation. It was observed that gonocyte

treatment with specific inhibitors for Raf1 (iRAF1) and for MEK1/2 (PD98059 and UO126)

significantly attenuated the proliferative response to PDGF-BB following overnight treatment,

while the PI3K inhibitor, wortmannin, did not have a negative effect on PDGF-induced

proliferation of gonocytes at PND3 (figure 7). In order to address the hypothesis that the

effects of PDGF and estrogen are interdependent and crosstalk to stimulate proliferation,

studies were performed where inhibitors for one pathway were incubated with the stimulator

of the other pathway. ICI 182780 was able to negatively affect the proliferative effects of

PDGF-stimulated gonocytes, while AG370 reduced the response to gonocytes to estrogen

stimulation. Similarly, the inhibition of Raf1 and MEK1/2 in the MAPK cascade blocked the

estrogen-induced proliferation (figure 7).

12.3 In Vitro exposure to xenoestrogens and phytoestrogens induce proliferation in

neonatal gonocytes

Earlier experiments in the lab demonstrated that, in the same manner as 17β-estradiol

induced proliferation in PND3 gonocytes, exogenous estrogenic compounds commonly found in

the environment such as genistein and BPA as well as the potent xenoestrogen DES caused

gonocyte proliferation in a dose dependant manner, although the phytoestrogen coumestrol

did not. Addition of estrogen receptor antagonists and PDGF inhibitors showed that these

65

PI3K

DAPI 2d Ab DAPI

Mek1 Erk1/2Raf1 AMH

mouse IgG

DAPI

Mer

geAb

Day 0 Day 1

rabbit IgG DAPI

Raf1 MEK1 ERK1/2 PI3K Non Specific

Hoec

hst

Spec

ific

AbM

erge

MIS/AMH2d Ab

mIgG

Rb IgG

tubulin

S/M

Gonocytes

Raf1

Mek1

Erk1/2

MW(kDa)

4442

72

55

43

4442

72

55

43

tubulin

Raf1

Mek1

Erk1/2

A

B

C

66

Figure 6: Identification of the expression of MAPK downstreamsignalling pathways, PI3K and MIS/AMH in PND3 isolatedgonocytes and testis sections. Previous studies in our lab identifiedmRNA levels of MAPK molecules. (A) Immunohistochemistry wasperformed on PND3 rat testis sections to observe protein profiles ofmolecules that are generally associated with PDGF signalling in thewhole testis. Specific fluorescent antibodies were used to detectprotein expression and a Hoechst nuclear dye was used to identifysingle cells in the tissue. Negative controls shown for non-specificIgG and secondary antibodies alone. (B) Isolated gonocytes werecollected on cytospin slides and the cells treated forimmunocytochemistry at day 0, right after the BSA gradient, andafter one day in culture. Negative controls are shown for rabbit andmouse non-specific IgG and secondary antibody alone. Nuclearstaining was performed using DAPI. (C) Western blot analysisconfirmed protein profiles, and was used to compare proteinexpression in gonocytes to Sertoli and Myoid cells (S/M). Thegonocytes were lysed on day 0, directly after the BSA gradient.

67

compounds behaved similarly to 17β-estradiol. However, in these experiments performed in

the presence of 2.5% serum, genistein and BPA acted at concentrations much lower than

anticipated from their affinity for ERs. Indeed, 17β-estradiol induced proliferation at

concentrations of 10-9M and above, while genistein, BPA and DES were all able to stimulate

proliferation starting at 10-12M. Thus we conducted further experimentation to explore

whether the presence of serum could explain these data. Dose response studies were

performed with genistein and BPA in media deficient of FBS, PDGF-depleted serum or with

charcoal-stripped FBS to see how the xenoestrogens respond in the absence of steroids or

growth factors from the serum. When no FBS was used, PDGF treatment required the addition

of an estrogenic compound such as 17β-estradiol, genistein or BPA to induce proliferation,

reaching a 2-3 fold increase at 10-6M for all of the estrogens (figure 8A). Gonocytes treated in

the presence of PDGF-depleted serum required both PDGF and 10-9M estrogenic compound in

order to induce the same 2-3 fold increase in proliferation. Significant proliferation was also

observed using 10-9M genistein, BPA or 17β-estradiol in the charcoal stripped FBS, which should

still contain other growth factors. These experiments showed that, under these conditions,

exogenous estrogens work in the same manner as 17β-estradiol with similar dose-response

curves (figure 8B). Given that there is little potential to be exposed to only one exogenous

estrogenic compound at a time, we also exposed gonocytes simultaneously to both genistein

and BPA, each in minute, fentomolar concentrations. We had observed that at these

concentrations, neither of the two molecules induced gonocyte proliferation. However, the

two molecules added together indeed did result in a significant increase in gonocyte

proliferation (figure 8C).

12.4 Treatment of Gonocytes with Other Steroid Hormones

As it is well documented that testosterone is the “male sex hormone”, we investigated

whether or not testosterone and another steroid hormone important in sexual differentiation

and function, progesterone, played a role in the proliferative stage of PND3 gonocytes.

Gonocytes were treated with 10-6M testosterone or 10-6M progesterone overnight with or

without FBS. Increased proliferation above the basal rate was not observed in these

68

***

**

PDGF

AG-3

70+ -

- ++ +

- -

Proliferation rate (%)40 30 20 10 0

**

PDGF

iRaf

1- -

+ -- +

+ +

Proliferation rate (%)40 30 20 10 0

**

**

PDGF

UO

126

PD98

059

- - -

+ - -

- + -

+ + -

- - +

+ - +

Proliferation rate (%) 40 30 20 10 0PD

GFW

ortm

anni

n

***

ns

- -+ -

- ++ +

Proliferation rate (%)40 30 20 10 0

*

E2iR

af1

D**

**

- -+ -

- ++ +

Proliferation rate (%)40 30 20 10 0E2

UO

126

PD98

059

***

**

- - -

+ - -

- + -

+ + -

- - +

+ - +

40 30 20 10 0

E

Proliferation rate (%)

***

***

PDGF E2

AG-3

70

- - -

+ - -

- + -

+ + -

- - +

- + +

***

40 30 20 10 050 Proliferation rate (%)

***

***

***

E2IC

I 182

780

- -+ -

- ++ +

Proliferation rate (%) 40 30 20 10 050*

***

***

PDGF

ICI 1

8278

0- -

+ -- +

Proliferation rate (%)40 30 20 10 0

+ +

***

Figu

re7:

PDG

Fan

des

trog

en-in

duce

dpr

olife

ratio

nof

gono

cyte

sin

the

pres

ence

ofsp

ecifi

cin

hibi

tors

.PN

D3go

nocy

tes

wer

etr

eate

dov

erni

ghtw

ithth

efo

llow

ing

trea

tmen

tsin

the

pres

ence

of2.

5%FB

San

dBr

dUin

cultu

reto

mea

sure

the

prol

ifera

tion

rate

ofth

ece

llsin

each

trea

tmen

t.Go

nocy

tes

wer

etr

eate

dw

ithPD

GFw

ithan

dw

ithou

tsp

ecifi

cin

hibi

tors

for

itsty

rosin

eki

nase

rece

ptor

(AG3

70)

and

dow

nstr

eam

path

way

,bl

ocki

ngRa

f1(iR

af1)

and

MEK

1(U

O12

6an

dPD

9805

9).

Incu

batio

nw

ith17

beta

-es

trad

iolw

aspe

rform

edw

ithor

with

outI

CI18

2780

.sim

ilarly

PDGF

trea

tmen

twas

incu

bate

dw

ithth

eER

anta

goni

stan

dce

llsw

ere

trea

ted

with

estr

ogen

alon

gsid

ein

hibi

tors

ofth

ePD

GFR

and

dow

nstr

eam

path

way

.The

sere

sults

indi

cate

acr

osst

alk

betw

een

the

two

path

way

s.Ce

llsw

ere

colle

cted

oncy

tosp

insli

des

and

coun

ted

forp

ositi

vepr

olife

ratio

nst

aini

ng.T

hepe

rcen

tofp

rolif

erat

ing

cells

show

nar

eth

em

eans

±se

mof

seve

rale

xper

imen

ts.

69

Figure 8: Effects of xenoestrogens on PND3 gonocytes in the absence of serum orpresence of PDGF-depleted FBS. Isolated gonocytes from PND3 rats were treated with 10ng/ml PDGF and various concentrations of E2, genistein (gen) and Bisphenol A (BPA)overnight in RPMI. (A) Dose dependant effect of estrogens in the presence of PDGF andno FBS. The two xeno/phyto-estrogens were incubated at different concentrations withPDGF to see if the exogenous estrogens behaved similarly to E2 in gonocytes. (B) BPA andgenistein were used to treat gonocytes in PDGF-depleted serum with or without PDGF toverify their requirement of PDGF in inducing proliferation. Proliferation rates werecalculated to determine which treatments induced more than the basal proliferation. (C)Proliferation studies were conducted comparing treatment of single estrogeniccompounds to two estrogens together. (D) Chemical structures of the tested exogenousestrogenic compounds. * P<0.05

B

Prol

ifera

tion

rate

(%)

10-12 10-9

Estrogen (M)

10-6

*

*

0

10

20

30

40

50

60

PDGF

Prol

ifera

tion

rate

(%)

17β-estradiol

GenisteinBPA

+---

----

---+

+-+-

--+-

+--+

-+--

++--

** *

PDGF-depleted FBSNo FBSA BPr

olife

ratio

n ra

te (%

) 50

40

30

20

10

0

**C

XenoestrogensDES

BPA

PhytoestrogensGenistein

Coumestrol

D

70

2.5% FBS No FBS

Control

PDGF+E2

Progesterone

TestosteronePr

olife

ratio

n ra

te (%

)

***

Figure 9: Steroid hormone proliferative effect on PND3 gonocytes with and without serum. PND3 gonocytes were treated with 10-6 M testosterone and progesterone and examined for proliferative effect using BrdU in the presence and absence of FBS.

71

experiments, which confirms that the effect of estrogenic compounds on gonocytes is specific,

and that gonocytes are not stimulated to proliferate by the other main steroids produced in the

testis (figure 9).

12.5 ERK2 Activation via PDGF-BB and 17β-estradiol

In order to decipher the molecules involved in the crosstalk between PDGF and estrogen

pathways in proliferative gonocytes, we performed short term time course experiments. We

investigated the phosphorylation of ERK1/2, a molecule that we had found abundantly

expressed in gonocytes, to examine whether activation of the signalling cascade leading to

transcription regulated by the ERK1/2 signaling molecule could serve as the common molecular

element and common target system of the PDGF-estrogen crosstalk. It should be noted that the

extremely low number of cells isolated in each experiment limited the numbers of conditions

that could be tested simultaneously, making it hard to compare each condition with its specific

inhibitor with the same isolation of cells. To appreciate the short term progression of the

pathway, cells were cultured for 15, 30 and 60 minutes in order to make isolated observations

at various short term time points, looking at the expression and phosphorylation (activation) of

ERK. Using Western blot analysis, we were able to distinguish between the two isoforms, ERK1

and ERK2, which are found at 44 and 42 kDa respectively. While the stimulation of gonocytes

with PDGF or estrogen for 15 min showed increased ERK2 phosphorylation in some

experiments, a consistent and overall significant increase of phosphorylated ERK2 was observed

only with the combination of PDGF and estrogen together in four independent experiments

(figure 10B). After 30 minutes of incubation, all 3 conditions caused a strong phosphorylation

of ERK2, while ERK1 phosphorylation remained significantly closer to the basal level (2

experiments). Interestingly, 15 min treatment with genistein, and to a lesser extent BPA,

induced a small increase in ERK2 phosphorylation, yet much lower than the combination of

PDGF + E2 (figure 10D). The fact that ERK2 is the main phosphorylated form of ERK is

interesting given our observation that PND3 gonocytes in vehicle alone appear to express a

much higher concentration of total ERK1 protein, rather than ERK2, and gonocytes treated with

PDGF-BB and 17β-estradiol for 60 minutes and 4 to 8 hours appear to increase their expression

72

of ERK1 protein, which remains in abundance, compared with ERK2 (figure 10B). A significant

increase in phosphorylated ERK2 was observed after 60 minutes stimulation by PDGF and

estrogen added together, but not by any of the single agents (4 independent experiments).

While the presence of UO126 inhibited the effect of PDGF + E2 on ERK2 phosphorylation, the

addition of AG370 did not block ERK2 activation in gonocytes (figure 10A).

Using immunocytochemistry, which could not distinguish between ERK1 and 2, one

could not clearly see the maintenance of phosphorylated ERK2 after 60 min that we had

measured by western blot analysis (figure 10F). However, manual counting of the number of

cells immunopositive for phospho-ERK versus negative cells, showed a clear increase at 60

minutes in the percentage of cells expressing high levels of phosphorylated ERK when treated

with PDGF, as well as a smaller increase observed when gonocytes were treated with PDGF and

E2 together (figure 10E).

The determination by western blot analysis of ERK1 and 2 activation levels in gonocyte

nuclear and cytosolic fractions showed that the majority of the increase in phospho-ERK2 was

localized in the cytosolic fraction of the cells treated with PDGF + E2, while there was no clear

change from the basal level observed in the nuclear fractions after 15 min treatment (figure

10C). This result was confirmed by immunocytochemistry and confocal microscopy, where the

majority of cells treated with PDGF + E2 presented increased phospho-ERK in their cytosol after

30 min incubation, although some cells expressed high levels of phospho-ERK1/2 in their nuclei

figure 10G). It should be noted that PND3 gonocytes are not synchronized and that any given

preparation of gonocytes may include different proportions of quiescent, mitotic and migratory

cells in function of the exact ages of the pups at the time of sacrifice (Culty, 2009). An

interesting observation is that under confocal microscopy it appears that some of the cells

expressing high levels of nuclear phospho-ERK were in the process of proliferation, being

observed in one of the later stages of mitosis (figure 14). Thus, it is possible that gonocyte

proliferation requires ERK translocation into the nucleus during one or more of the phases of

mitosis.

73

30 M

inut

e G

onoc

yte

Trea

tmen

tPh

osph

oryl

ated

ERK

CP

EP+E

P+E+A

G P+E+U

O

P+E+IC

I0246810

Trea

tmen

ts

pERK expression(fold change / Control)

15 M

inut

e G

onoc

yte

Trea

tmen

tPh

osph

oryl

ated

ERK

CP

EP+E

P+E+A

G+UO+IC

I01234

Trea

tmen

ts

pERK expression(fold change / Control)

**

60 M

inut

e G

onoc

yte

Trea

tmen

tPh

osph

oryl

ated

ERK

CP

EP+E P+E

+AG P+E

+UO P+E

+ICI

P+E+A

G+UO+IC

I012351015

Trea

tmen

ts

pERK expression(fold change / Control)

*

*

4 ho

ur tr

eatm

ent

Cont

rol

P+E

PDGF

Estr

ogen

Cont

rol

P+E

PDGF

Estr

ogen

Post

Nat

al d

ay 3

8 ho

ur tr

eatm

ent

Hist

one

P-ER

K

ERK

Phos

phor

ylat

ed E

RK1/

2N

ucle

ar E

xtra

ctio

n Ex

perim

ent

44 42kDa

Nuc

leus

Cyto

sol

CC

CC

P+E

P+E

P+E

P+E

15 m

inut

es in

trea

tmen

t60

min

utes

in tr

eatm

ent

ERK

pERK

Tubu

lin

Cont

rol

PDGF

Estr

ogen

P+E

Cont

rol

PDGF

Estr

ogen

P+E

Post

Nat

al D

ay 3

CP+

EG

BPA

pERK ERK

Tubu

lin

PND3

15

min

ute

trea

tmen

t

A BC D

74

Imm

unoc

ytoc

hem

istr

y30

Min

ute

Gon

ocyt

e Tr

eatm

ent

phos

phor

ylat

ed E

RK

CP

EP+E

AG+UO+IC

I P+AG P+U

O E+ICI

P+E+A

G P+E+U

O P+E+IC

I

P+E+A

LL020406080100

Trea

tmen

ts

Percentage of cells expressing phosphorylated ERK

Imm

unoc

ytoc

hem

istr

y60

Min

ute

Gon

ocyt

e Tr

eatm

ent

phos

phor

ylat

ed E

RK

CP

EP+E

AG+UO+IC

I P+AG P+U

O E+ICI

P+E+A

G P+E+U

O P+E+IC

I

P+E+A

LL020406080

Trea

tmen

ts

Percentage of cells expressing phosphorylated ERK

*

Imm

unoc

ytoc

hem

istr

y30

Min

ute

Gon

ocyt

e Tr

eatm

ent

phos

phor

ylat

ed E

RK in

the

nucl

eus

CP

EP+E

AG+UO+IC

I P+AG P+U

O E+ICI

P+E+A

G P+E+U

O P+E+IC

I

P+E+A

G+UO+IC

I0.

0

0.5

1.0

1.5

2.0

Trea

tmen

ts

Expression of pERK / control

Imm

unoc

ytoc

hem

istr

y30

Min

ute

Gon

ocyt

e Tr

eatm

ent

phos

phor

ylat

ed E

RK in

the

cyto

sol

CP

EP+E

AG+UO+IC

I P+AG P+U

O E+ICI

P+E+A

G P+E+U

O P+E+IC

I

P+E+A

G+UO+IC

I0.

0

0.5

1.0

1.5

2.0

2.5

Trea

tmen

ts

Expression of pERK / control

P+EP EC

pERK

30pE

RK60

pERK

30C

P

EP+

E

cytosolicstaining

E F

G

75

Figure 10: PDGF and estrogen-induced Activation of ERK2 in short term assay.PND3 isolated gonocytes were treated with PDGF with or without estrogen andinhibitors for 15, 30 and 60 minutes. Cells were stopped by either centrifugationor paraformaldehyde and then lysed for western blot analysis or fixed to cells viacytospin centrifugation for immunocytochemistry experiments. (A) western blotprofiles of phosphorylated ERK2. the gels were evenly loaded with the samequantities of protein per well, and therefore we normalized the densitometry ofthe protein bands on the membrane only to the control value of eachexperiment. The results are expressed as the changes in ERK activation betweencontrol (1) and treatment. (B) western blot representations of ERK, phospho-ERKand tubulin/Histone for PND3 gonocytes treated for 15 and 60 minutes as well as4 and 8 hours. (C) Representative membrane expressing phosphorylated ERKfrom a nuclear extraction experiment where the nuclear and cytosolic fractionswere separated and investigated for differences in protein profiles. ERK activationappears to be more prominent in the cytosol than the nucleus. (D) As we havealready observed that genistein and BPA modulate gonocyte proliferation in asimilar manner to 17β-estradiol, this western blot representation shows howgenistein also activates ERK in short term cultures. (E) immunocytochemistryexperiments were performed for short term ERK activation. Pictures of the cellswere captured and cells were counted for activation or inactivation. The data isrepresented as a percentage of the total population of activated cells. (F) Totaland nuclear intensity of immunoreactivity of ICC were calculated. Subtraction ofthe nuclear fraction from the total cell gave the intensity of the given protein inthe cytosolic ring. These histograms compare changes in the control and treatedcell expression of phospho-ERK between the nucleus and cytosol. (G)representative ICC fluorescent pictures of control and treated cells colocalizedwith DAPI nuclear stain. The bottom four pictures are the primary anti phospho-ERK immunoreactivity without the DAPI showing that most of the cells have adistinct cytosolic rings.

76

As expected, UO126 inhibited the ERK1/2 activation in gonocytes treated with PDGF.

Surprisingly, ICI together with E2 or added to the combination of PDGF + E2 induced a stronger

phospho-ERK immunoreactive signal in gonocytes and a higher ratio of positive cells than when

E2 or PDGF were added alone, despite the long term inhibitory effect exerted by ICI on E2- and

PDGF-induced proliferation (figure 10E, figure 11). The same observation was made by western

blot analysis of cells exposed to ICI (figure 10A).

12.6 PDGF and Estrogen Increase Expression of PDGFRβ and ERβ Immunoreactivity

In Vitro

Treatment of PND3 gonocytes with PDGF-BB induced a significant increase in PDGFRβ

immunoreactivity in the cells in general and specifically in the cytosol, as seen via

immunocytochemistry. This expression was seen through the quantification of

immunofluorescent signals in treated cells using an anti PDGFRβ antibody that showed an

increase in PDGFRβ after 30 minutes of treatment with PDGF, estrogen and the two in

combination (figure 12A). Confocal microscopy as well as immunocytochemistry confirmed the

quantitative data, showing that a majority of cells expressed PDGFRβ in the cytosol (figure 12B).

PDGFRβ immunofluorescence was present at high amounts in several dividing cells, and

confocal microscopy analysis showed a number of cells expressing tremendous amounts of

PDGFRβ in the nucleus (figure 14). This is a clear example of the heterogeneity of the isolated

gonocyte population. The clarification of the different PDGFR identities and profiles could be

instrumental in understanding this critical stage of gonocyte development. The nuclear

PDGFRβ is very likely to be the V1 isoform as we observed in live cell imaging experiments (see

section on live cell imaging and discussion). Another important observation is that not all

gonocytes at this stage express PDGFRβ and some cells were seen by confocal microscopy to be

negative for PDGFRβ expression.

Unlike PDGFRβ and activated ERK, which are most commonly found in the cytosol of

gonocytes, ERβ has a number of different expression profiles, all of which are expressed

throughout early neonatal life. Confocal microscopic imaging of PND3 gonocytes stained with

an ERβ antibody presented up to four different protein profiles in a single section. Gonocytes

77

were found to either be completely negative for ERβ, expressed ERβ mostly in the cytosol or

mostly in the nucleus, and on rare occasions, expressed ERβ in the nucleoli of the cell (figure

14). However, ICC analysis showed that the majority of ERβ-expressing gonocytes exhibited

simultaneously cytosolic and nuclear signals (figure 13). Gonocytes with pseudopod-like

projections often expressed high levels of ERβ in the pseudopod (figure 14). An interesting

observation could be made by examining protein expression in dividing cells. Noticeably, some

dividing cells expressed high ERβ levels while others did not. Understanding at which phases

ERβ is or is not expressed can be very important to our understanding of its role in gonocyte

proliferation.

Treatment with PDGF-BB or 17β-estradiol for 60 minutes induced an increase in ERβ

immunoreactivity in the cytosolic ring. These data were supported by western blot analysis of

nuclear and cytosolic fractions of gonocytes in which 60 minutes treatment with either PDGF-

BB or 17β-estradiol caused an apparent increase in cytosolic ERβ compared to the control

(figure 13).

12.7 Live Cell Imaging of Gonocytes Transfected with an EGFP-V1-PDGFRβ

Construct, Preliminary Observations

Because the ICC data demonstrated a relatively strong signal for PDGFRβ in the cytosol

and nucleus that most likely corresponded to the V1-PDGFRβ variant previously found in

gonocytes, preliminary experiments were conducted to investigate its expression and

subcellular localization in control versus PDGF and E2 treated cells. Our laboratory previously

showed that V1-PDGFRβ is missing part of the transmembrane domain, and it was found to be

expressed in the cytosol of transfected cell lines (Wang and Culty, 2007). A first attempt to

transfect PND3 gonocytes for 24 hours without serum yielded only 10-20% positive cells.

Because gonocytes represent roughly 1% of the total cell population isolated from the

seminiferous tubules of PND3 testes and thus the number of cells is always limiting, it was

critical to optimize the method and improve the yield of EGFP-V1-PDGFRβ transfected

gonocytes. We decided to dissect one day earlier, at PND2, in order to allow an extra day of

transfection before the BSA gradient, which separates gonocytes from the Sertoli cells, and to

78

ImmunocytochemistryGonocyte 30 minute treatment

phosphorylated ERKICI 182780 dose response study

CICI50

ICI100

E+ICI50

E+ICI100

0

20

40

60

Treatments

Perc

enta

ge o

f cel

ls e

xpre

ssin

g p

hosp

hory

late

d ER

K

Figure 11: Dose response curve for ICI treatment of gonocytes. A singleexperiment was performed to see if there was a dose dependent effect of ICI onthe cells. Immunocytocytochemistry was performed and fluorescent pictures weretaken of the cells. Cells expressing high levels of pERK were counted versus thecells that had barely any immunoreactive signal.Nearly a doubling of cells expressing phosphorylated ERK was seen when the ICIconcentration in gonocyte treatment for 30 minutes was doubled, reflecting ashort-term agonistic effect of ICI 182780 leading to ERK activation in gonocytes.

79

CP

EP+

EPD

GFR-

beta

30

min

utes

Imm

unoc

ytoc

hem

istr

y30

Min

ute

Gon

ocyt

e Tr

eatm

ent

PDG

FR-b

eta

in th

e nu

cleu

s

CP

EP+E

P+E+A

G P+E+U

O

P+E+IC

I0.

0

0.5

1.0

1.5

2.0

2.5

Trea

tmen

tsExpression of

PDGFR-beta / control

Imm

unoc

ytoc

hem

istr

y30

Min

ute

Gon

ocyt

e Tr

eatm

ent

PDG

FR-b

eta

in th

e w

hole

cel

l

CP

EP+E

P+E+A

G P+E+U

O

P+E+IC

I0.

0

0.5

1.0

1.5

2.0

Trea

tmen

ts

Expression ofPDGFR-beta / control

Imm

unoc

ytoc

hem

istr

y30

Min

ute

Gon

ocyt

e Tr

eatm

ent

PDG

FR-b

eta

in th

e cy

toso

l

CP

EP+E

P+E+A

G P+E+U

O

P+E+IC

I0.

0

0.5

1.0

1.5

2.0

Trea

tmen

ts

Expression ofPDGFR-beta / control

**

*

A B

80

Figure 12: Effects of PDGF and estrogen treatment on PDGFR-betain short term PND3 gonocyte assays. (A) ICC quantification ofimmunofluorescent intensity was calculated as a total intensity ofthe signal for each cell. Quantifications were taken for the nucleararea and the whole cell and subtraction of the two provided theintensity in the cytosolic ring. Histograms show the changes in thetreatments against the control values for the nucleus, cytosol andtotal cell intensities. Changes in the immunofluorescence profiles ofPDGFR-beta could be seen at 30 minutes in culture. (B)Representative ICC pictures are shown for control and treated cellsfor 30 minutes in treatment. The top row shows the PDGFR-betaprotein in each treatment, while the bottom row colocalizes theprotein with the nuclear DAPI stain. * p< 0.05

81

Imm

unoc

ytoc

hem

istr

y60

Min

ute

Gon

ocyt

e Tr

eatm

ent

estr

ogen

rece

ptor

bet

a in

the

cyto

sol

CP

EP+E

E+ICI P+E

+AG P+E

+UO P+E

+ICI

P+E+A

G+UO+IC

I0.

0

0.5

1.0

1.5

2.0

2.5

Trea

tmen

ts

Expression of ER-beta / control

Imm

unoc

ytoc

hem

istr

y60

Min

ute

Gon

ocyt

e Tr

eatm

ent

estr

ogen

rece

ptor

bet

a in

the

nucl

eus

CP

EP+E

E+ICI P+E

+AG P+E

+UO P+E

+ICI

P+E+A

G+UO+IC

I0.

0

0.5

1.0

1.5

2.0

Trea

tmen

tsExpression of ER-beta / control

Imm

unoc

ytoc

hem

istr

y60

Min

ute

Gon

ocyt

e Tr

eatm

ent

estr

ogen

rece

ptor

bet

a in

the

who

le c

ell

CP

EP+E

E+ICI P+E

+AG P+E

+UO P+E

+ICI

P+E+A

G+UO+IC

I0.

0

0.5

1.0

1.5

2.0

Trea

tmen

ts

Expression ofER-beta / control

*

CP

EP+

EER

-bet

a 60

min

utes

ER-b

eta

spec

ific

pept

ide

A B

82

Figure 13: Effects of PDGF and estrogen treatment on estrogen receptor-beta in short term PND3 gonocyte assays. (A) Histograms represented in thesame fashion as in the previous figure. Changes in the immunofluorescenceprofiles of ER-beta could be seen at 60 minutes in culture. (B) RepresentativeICC pictures are shown for control and treated cells for 60 minutes intreatment. The top row shows the ER-beta protein in each treatment, whilethe bottom row colocalizes the protein with the nuclear DAPI stain. Negativecontrols were conducted with primary antibody being pre-incubated with itsspecific peptide. * p< 0.05

83

p-ER

KPD

GFR

βER

β

84

ERβ

PDGFRβ p-ERK

Proliferating gonocytes

85

Figure 14: Representative gonocyte images captured with confocal microscopy forlocalization of phosphorylated ERK (p-ERK), PDGFRβ and ERβ. (A) PND3 gonocyteswere incubated for 30 to 60 minutes with various combinations of PDGF and 17β-estradiol. Due to the time constraints and low number of cells when dealing withprimary cultures and confocal microscopy, at this point we have not observedsteady trends between the different treatments. However, by carefully analyzing alarge population of cells under the confocal microscope, we can make a number ofimportant observations that can provide some insights on possible mechanismsinvolved in the different phases of neonatal gonocyte development. ERβ is found inmany different patterns in gonocytes and there is not one protein profile thatstands out as the most common. Some cells have a very distinct cytosolic ring whileothers express ERβ only in the nucleus and others only in the nucleoli. PDGFRβ ispreferentially expressed in the cytosol with most cells presenting a very distinct ringaround the nucleus, while a smaller sub-population of gonocytes do not expressPDGFRβ at all. However, there is also a distinct subset of gonocytes that expressPDGFRβ in the nucleus, including dividing gonocytes, suggesting that a nuclear formof PDGFRβ might be involved in mitosis. p-ERK is also localized mainly in the cytosolof p-ERK positive gonocytes, while other cells were completely negative, showingabsence of activated ERK in these gonocytes. (B) Thanks to a few cells that werestopped during proliferation, we were able to observe potential differences inprotein profiles of actively dividing cells versus non-proliferating cells. Again ERβshowed different profiles which could be due to different phases of mitosis. Most ofthe cells did express ERβ in the nucleus, and only one cell had very spotty staining.PDGFRβ clearly localized in distinct areas of the nucleus during proliferation, butalso in the cytosol. p-ERK was also highly expressed in dividing cells and probablylocated in the nucleus.

86

Cont

rol

PDGF

17β-

Estr

adio

l

87

P+E

AG+U

O+I

CIP+

E+AG

+OU

+ICI

88

Figure 15: Live Cell Imaging of GFP bound V1-PDGFRβ transfected into PND3gonocytes. Cells were transfected for 2 days with a mamalian expression vectorcontaining a GFP-V1-PDGFRβ insert and they were then treated for 60 minuteswith PDGF, estrogen and/or inhibitors. Protein expression profiles wereobserved in order to examine whether PDGF or estrogen induce changes inexpression or subcellular localization of the variant receptor. As we onlyperformed this experiment one time in full, we did not obtain alarge enough number of cells per treatment that would allow finding specifictrends. The V1 variant appeared to be expressed in the nuclei of most cells,although it is interesting that with P+E+AG+UO+ICI, it was seen in a ring aroundthe nucleus, localized in the cytoplasm only.

89

get V1-PDGFRβ transcript and protein expression in the gonocytes at the time of their

maturation to a mitotic phenotype (corresponding to PND3 and 4). After a second night of

transfection, the gonocytes were treated with different combinations of PDGF, Estradiol and

the inhibitors AG370, UO126 and ICI 182780 and observed live using the confocal microscope

for up to 60 minutes. Although these experiments were preliminary, a few observations could

be made from the positive gonocytes in culture. V1-PDGFRβ was primarily expressed in the

nucleus or in the whole cell, but more of the cells appeared to present a nuclear signal for V1-

PDGFRβ. However, in gonocytes incubated with inhibitors, V1-PDGFRβ appeared to be

localized in many cells in a visible cytosolic ring. We also noted that treatment with PDGF and

PDGF + E2 showed a higher expression of EGFP-V1-PDGFRβ than in the control and cells treated

with inhibitors. These preliminary results suggest that PDGF and estrogen may rapidly induce

the translation of V1-PDGFRβ mRNA and may be involved in its nuclear localization (figure 15).

13. Discussion

Because they are the precursor cells of the spermatogonial stem cells, gonocytes represent

a critical developmental stage of the germ line for spermatogenesis and male fertility.

Moreover, it is believed that they are a potential origin of testicular germ cell tumours (Rajpert-

De Meyts E. and Hoei-Hansen, 2007). Thus, understanding gonocyte development can provide

invaluable information in the battle against testicular cancer and its prevention. The goal of my

study was to gain insight on the mechanisms regulating gonocyte proliferation. Through in vitro

studies of purified PND3 rat gonocytes, we concluded that gonocyte proliferation is not only

stimulated by both PDGF and estrogen through their specific receptors, but that they are both

working interdependently. This was shown in studies where gonocytes were treated with PDGF

in the presence of the ER antagonist ICI 182780, and PDGF-induced proliferation was

attenuated. Similar findings were shown for estrogen-induced proliferation in the presence of

PDGFR inhibitors. In order to understand why treatment with only one of the two factors did

induce proliferation, gonocytes were treated with PDGF in charcoal-stripped, steroid depleted,

FBS, or with 17β-estradiol in serum devoid of PDGF. In both cases, induction of proliferation

90

was no longer observed, confirming the hypothesis that proliferation with either PDGF or 17β-

estradiol alone was supported by small amounts of the other factor present in FBS. In addition,

when molecules of the MAPK pathway were inhibited, estrogen-induced proliferation was no

longer detected. We therefore concluded that gonocyte proliferation at PND3 is induced

through crosstalk between estrogen and PDGF, acting through their specific receptors, although

the mechanism by which these two pathways communicated was yet to be defined. It was our

goal to investigate at which point of the receptor signalling and downstream cascades this

crosstalk takes place.

Our first step in investigating the direct effect of PDGF and estrogen on gonocyte

proliferation was to ensure that no external growth factors or steroids found in culture medium

or serum were interfering with our treatments, thereby providing stimulation above the desired

treatments. Experiments employing charcoal-stripped serum and PDGF-depleted serum

confirmed that both PDGF and estrogen together are required, in the absence of other factors,

to induce proliferation of PND3 gonocytes.

Interestingly, we found that exogenous estrogens such as BPA and genistein were able

to stimulate gonocyte proliferation in the absence of FBS but required the presence of PDGF

with a similar dose response curve to E2. This is in contrast to results obtained using normal FBS

where the exogenous estrogens acted at lower concentrations than E2. As well, exogenous

estrogens behaved similarly to E2 as their responses were also inhibited by the PDGFR inhibitor

AG370. The differences between the observations with and without serum could be due to the

presence of low PDGF and E2 levels where the background PDGF and estrogen might have

placed the cells in a more responsive state to the effects of exogenous estrogens. Moreover,

the amount of E2 bound to the steroid binding proteins from serum should be higher than the

amount of exogenous estrogens due to lower affinity of these compounds for steroid binding

proteins, leading to lower levels of free E2 as compared to the levels of exogenous estrogens.

This would result in higher levels of free exogenous estrogens available to stimulate ER in

comparison to E2 for a given concentration of added steroids. However, in the absence of

binding proteins from serum, all of the estrogenic compounds would present the same

91

amounts of free compound and thus the same proliferation efficacy. It is also possible that

different estrogenic compounds possess distinct binding affinities to ER and are able to

stimulate or be stimulated by other cofactors in their immediate environment. The presence of

serum could provide the necessary cofactors for exogenous estrogenic compounds to provide a

stronger response than E2 while without serum they both behave in the same way.

We found that testosterone and progesterone do not induce gonocyte proliferation.

This is a very important observation as the majority of reproductive development in males is

regulated in some way by testosterone, while progesterone and estrogen pathways frequently

interact in cells. Moreover, a recent study reported that foetal gonocytes express the androgen

receptor and that their numbers were decreased by androgen treatment. Our results suggest

that this is not the case in neonatal gonocytes and that the direct effect of androgen on germ

cells might be limited to the foetal period. Although testosterone levels are not as high in

neonatal testis as during foetal life or after puberty, neonatal Leydig cells secrete measurable

amount of testosterone (Culty et al, 2008). However, part of the testosterone produced at this

time is converted to estrogen by the aromatase present in Sertoli cells and to

dihydrotestosterone by the 5α-reductase to support the development of male reproductive

system. Testosterone may play an indirect role in gonocyte development via its action on

somatic cells. Nevertheless, through these observations, it appears that gonocyte proliferation

is highly specific and requires both PDGF and an estrogen for activation.

Gonocyte proliferation is a highly regulated process which is activated at around PND3

in rats and PND1.5 in mice, simultaneously with migration of the seminiferous cords to the

basement membrane. Only gonocytes that reach the basement membrane will further

differentiate into spermatogonia. Although one could assume that such precisely regulated

events would require specific sets of activators unique to either proliferation, migration or

differentiation, studies have shown that some pathways may be involved in more than one

function, such as shown by the dual role of PDGFRs in gonocyte proliferation and migration (Li

et al, 1997;Basciani et al, 2008). Not only do gonocytes depend on the Sertoli cells to produce

the factors regulating their development, but their responsiveness to these molecules is likely

92

regulated by Sertoli cells too, as shown by our laboratory’s earlier studies where quiescent

PND2 rat gonocytes became responsive to mitotic agents after an additional day of maturation

with Sertoli cells (Li et al, 1997). The identification of the MAPK and PI3K signalling pathways in

PND3 gonocytes was an important observation leading to a better understanding of normal

gonocyte development. Since ERK1/2, Raf1, Mek1, as well as PI3K were present in gonocytes,

we investigated which of the two pathways was involved in proliferation through the use of

specific inhibitors for Raf1 (iRAF1) and Mek1/2 (PD98059, UO126), as well as PI3K

(Wortmannin). These experiments showed that gonocyte proliferation is induced primarily

through the MAPK pathway. While PI3K was not required for proliferation, its inhibition led to

small increases in proliferation above the basal level, suggesting that PI3K might be required to

maintain quiescence in neonatal rats (Thuillier et al, 2010).

These experiments were conducted in enriched gonocyte cultures in order to observe

direct effects of PDGF and estrogen on gonocyte proliferation, as opposed to co-cultures or

organ cultures where the gonocytes are in communication with all of the somatic cells and

respond to the factors produced by the Sertoli cells, PMCs and Leydig cells. Although these

latter types of cultures are more similar to the in vivo conditions of gonocyte development,

they would not allow one to study the direct effects of PDGF, estrogen, MAPK and PI3K

inhibitors on gonocytes since most of these proteins are also present in Sertoli and myoid cells.

A critical observation of our study was that MAPK inhibitors were able to attenuate the

proliferative response of gonocytes to estrogen stimulation. This data suggested that ERK1/2 is

involved in the crosstalk mechanism between PDGF and estrogen pathways and its activation is

necessary for gonocyte proliferation. With these data in hand, we were able to continue to the

next stage of the study, which was to examine the kinetics of ERK1/2 activation and whether it

required only PDGF or both PDGF and E2. To this end, we performed short term treatments of

gonocytes and followed ERK1/2 activation by immunoblot analysis and immunocytochemistry

(ICC) analysis.

Western blot analysis showed that ERK2 but not ERK1 was phosphorylated rapidly by

both PDGF and estrogen although ERK1 was the more abundant of the two ERK forms, and that

93

long term ERK2 phosphorylation could only be sustained in the presence of both PDGF and

estrogen. Moreover, the fact that AG370 alone could not block ERK2 activation in cells treated

with PDGF and E2 suggested that ERK2 phosphorylation is not the direct result of PDGFR

activation, but rather it is phosphorylated via a mechanism that is common to both the

estrogenic and PDGF pathways, requiring the activation of both the PDGF and estrogen

pathways for this phosphorylation to take place. Indeed, the finding that xenoestrogens were

able to induce short-term ERK2 phosphorylation but at much lower levels than the combination

of PDGF and E2, while they mimicked the E2 proliferative effect when combined with PDGF,

further supported the idea that estrogens act simultaneously with PDGF to activate ERK2. The

same conclusion could be reached from our ICC data where UO126 but not AG370 was found to

block ERK phosphorylation by PDGF and estrogen. Since the presence of both factors is also

necessary for gonocyte proliferation, this identifies ERK2 phosphorylation as a central event in

the crosstalk of PDGF and estrogen.

It is interesting to note that the addition of ICI 182780 together with E2 or PDGF + E2

surprisingly increased phosphorylated ERK in short term assays. This was unpredicted given the

fact that the same inhibitor was shown to decrease the proliferative capabilities of PDGF and

estrogen in PND3 gonocytes cultured for 1 day in medium containing 2.5% FBS. Although ICI is

a well known estrogen antagonist, there is sufficient documentation indicating its agonistic

effect in specific tissues. ICI 182780 stimulation actually induces MAPK signalling in specific

cells (Zhao et al, 2006;Mercier et al, 2003). Our results suggest that ICI 182780 can induce an

agonistic rapid non-genomic response, while behaving as an estrogen receptor antagonist in

long term, “classic” nuclear effects of ERs. While the existence of non-classic membrane bound

ERs such as GPR30 (Maggiolini and Picard, 2010) or alternative estrogen receptors such as the

ERRs have been reported in male reproductive tissues (Lazari et al, 2009), we have not yet

investigated their presence in PND3 gonocytes. A recent article showed that ERα might play an

important role in early germ cell development, and that ERα actually has a non-genomic effect

in PGCs, as well in the development of foetal germ cells between 11.5 and 12.5 dpc in mice (La

Sala et al, 2010). Indeed, our own studies had shown that besides a very large levels of ERβ,

94

one could detect a weak immunoreactive band of ERα in PND3 gonocytes (Wang et al, 2004),

suggesting that ERα might play a role in neonatal gonocytes.

This study also revealed changes in the levels and sub-cellular localization of ERβ and

PDGFRβ occur within an hour of stimulation with PDGF and E2, suggesting a complex and

dynamic interplay between the receptors in gonocytes. The sub-cellular localization of PDGFRβ

beyond its classical plasma membrane location suggests that the increased expression

corresponds to the induction of V1-PDGFRβ mRNA translation in gonocytes, although one

cannot exclude that it could reflect the internalization and turnover of full length PDGFRs after

their activation.

In view of these results, I propose that this crosstalk is more complicated than a simple dual

activation of MAPK, and that the proliferation process might be regulated at numerous levels.

While ERK activation is important, it appears that it is not the only function of PDGFR activation,

which is required but not sufficient to induce proliferation. Similarly, ERK does not exclusively

require PDGFR for its activation, as it is possible that the rapid, non-genomic response of

estrogen activates ERK, as it was reported for other cell types (Bouskine et al, 2008). There are

many possible mechanisms for this crosstalk given our results. A) PDGF via PDGFRs and E2 via a

non-classical ER could both activate different pools of ERK2 which would then activate the

expression of distinct genes required for proliferation. B) While only one, either PDGF or E2

could induce ERK2 phosphorylation, the other would protect phospho-ERK2 from

dephosphorylation by activating an unknown protein supporting the long term maintenance of

phospho-ERK. C) E2 effects could involve the activation and /or delocalization of both classical

and non-classical ERs. D) One of PDGF’s roles may be to activate V1-PDGFRβ expression and/or

delocalization via ERK activation as a negative feedback signal to cease proliferation. E) AG370

could inhibit V1-PDGFRβ besides the full length receptors, complicating the interpretation of its

effects. F) It is also possible that PDGFR and ERK are required for activation of nuclear ERs and

although with AG370 or ICI 182780 ERK can be phosphorylated by either mER or another

mechanism, PDGFR will not be available to activate nuclear ER (with AG370) or the ER will be

inhibited (by ICI 182780).

95

In support of these hypotheses, we know that gonocytes treated with PDGF have an

increased expression of PDGFRβ, while 30 minute treatment with estrogen increases the

protein expression of ERβ in the cytosol. The PDGFRβ increased in this system is most probably

that of the V1-PDGFRβ variant form. Indeed, live cell imaging data of EGFP-V1-PDGFRβ

expressing cells showed that the presence of PDGFR, ER and MEK inhibitors with PDGF and

estrogen led to V1-PDGFRβ being localized in the cytosol, in contrast with its nuclear expression

in the absence of inhibitors. This suggests that PDGF and estrogen are required for

maintenance of the functional activity and localization of V1-PDGFRβ, although their role in

neonatal gonocytes has yet to be defined. It is less probable that it reflects a recycling

mechanism of the membrane bound receptor, given that the confocal localization of PDGFRβ

varies; in some cells it is found in the nucleus, and other cells it is expressed at high levels of in

pseudopods. Observations of dividing gonocytes by confocal microscopy could possibly lead us

closer to achieving our goal of understanding the progression from quiescent to proliferative

and migratory gonocytes and then to SSCs and differentiating spermatogonia. If we can

understand which proteins must be expressed or activated during the different phases of

mitosis, we should gain a greater appreciation of the dynamic changes occurring in PND3

gonocytes. Given the heterogeneity of the cell population at PND3, as the timeframe between

different periods in neonatal gonocytes are not synchronized and can be as short as a few

hours, finding a way to identify the different sub-sets of gonocytes could prove to be very

important. From our experiments, it appears that gonocytes express high levels of phospho-

ERK during mitosis, either in their nuclei or in the cytosol. They also express ERβ in the nucleus

in some stages and in the cytosol at other times, and express intracellular forms of PDGFRβ,

most probably, V1- PDGFRβ.

In future experiments, it will be interesting to label gonocytes with BrdU or to measure

PCNA expression in order to identify proliferating cells, simultaneously with the determination

of ERK activation. In this way, one could know whether or not a cell is pre- or post-mitotic in

relation to the activation of signalling molecules in short term experiments, including but not

limited to ERK. It is clear that there are likely more than two distinct populations of gonocytes

at PND3, but it is important to at least understand what happens to the molecules that are

96

necessary for proliferation to occur. Moreover, it is unclear whether or not the cells undergo

only one mitosis and then arrest in preparation for differentiation, or if they continue to

proliferate until they are stimulated for differentiation. Looking at single cells is very

complicated in that so far there is no way to tell which stage the cells are in, except when clear

mitotic figures are observed. However, confocal microscopy using multiple antibodies

recognizing ERK, ERβ, PDGFRβ, alternative estrogen receptors and other signalling molecules

could provide important data by revealing simultaneous or successive changes in specific

protein levels and sub-cellular localizations in response to various treatments and time-courses.

In conclusion, my work uncovered new aspects of the regulation of gonocyte

proliferation, highlighting the critical role of ERK2 activation in this process, but also the fact

that ERK phosphorylation is under the dual control of PDGF and estrogen. Although the short-

time studies did not entirely identified the sequence of events occurring downstream of PDGF

and estrogen stimulation, it revealed the existence of a complex interplay between ERK

activation and the levels and sub-cellular localization of PDGF and estrogen receptors.

Moreover, this study showed that estrogen exerts rapid non-genomic effects on gonocytes and

that ICI 182780 behaved as an ER antagonist for long term incubations but as an agonist in

short-term stimulation, suggesting that there may be more than one species of estrogen

receptors in gonocytes. Finally, taken together with earlier studies in our laboratory, changes in

the sub-cellular localization of PDGFRβ suggested that the variant V1- PDGFRβ might play a role

in gonocyte proliferation or in its interruption to allow the further development of the cells.

97

Gon

ocyt

e

PRO

LIFE

RATI

ON

PDG

F

EE

R

E

Gen

e ex

pres

sion

?

Raf

ME

K

EE

RE

ER

PDG

F

ER

K

ER

K

p

Em

ER R

af

ME

K

ER

Kp

V1

UO

126

AG37

0

Tran

slatio

n Ac

tivat

ion

Expr

essio

n

Expr

essio

n

ICI

??

pPD

GF

?

98

Figure 16: Proposed Mechanisms of PDGF and Estrogen Crosstalk in GonocyteProliferation. In light of our observations that UO126 attenuates the phosphorylatedERK signal but AG370 does not and ICI 182780 actually increases ERK phosphorylationin short term treatments, we proposed possible mechanisms in our discussion as tohow PDGF and estrogen-induced proliferation through MAPK might occur. We proposethat short term rapid ER signalling is through a membrane receptor or another type ofnon-genomic mechanism, and this is most likely an important factor in ERKphosphorylation. It is possible that while ICI blocks the nuclear receptor, it works as anagonist for the membrane receptor. We also observed that P and E treatmentsincrease PDGFRβ and ERβ immunofluorescence. We proposed that the increasedPDGFR β could be an induction of translation of the variant V1 form that could play arole in the desensitization of the cells to PDGF to allow differentiation to proceed. Thiscould work through ERK activation via ER or membrane bound PDGFR. However, in theabsence of corresponding western blot analysis, we cannot exclude that the changesin intensity might reflect unmasking of an existing receptor, either by conformationalchange or by change in sub-cellular localization. Similarly, ERK activation might berequired to come from both ER and PDGFR to induce proliferation, but shutting off oneof these receptors might not be enough to attenuate ERK signalling.

99

Literature Cited

Alberts B, Johnson A, Lewis J, Raff M, Roberts K, Walter P (2002) Molecular Biology of the Cell. New York:Garland Science. 1463 p.

Anderson GL, Limacher M, Assaf AR, Bassford T, Beresford SA, Black H, Bonds D, Brunner R, Brzyski R, Caan B, Chlebowski R, Curb D, Gass M, Hays J, Heiss G, Hendrix S, Howard BV, Hsia J, Hubbell A, Jackson R, Johnson KC, Judd H, Kotchen JM, Kuller L, LaCroix AZ, Lane D, Langer RD, Lasser N, Lewis CE, Manson J, Margolis K, Ockene J, O'Sullivan MJ, Phillips L, Prentice RL, Ritenbaugh C, Robbins J, Rossouw JE, Sarto G, Stefanick ML, Van HL, Wactawski-Wende J, Wallace R, Wassertheil-Smoller S (2004). Effects of conjugated equine estrogen in postmenopausal women with hysterectomy: the Women's Health Initiative randomized controlled trial. JAMA 291:1701-1712.

Ashman LK, Cambareri AC, To LB, Levinsky RJ, Juttner CA (1991). Expression of the YB5.B8 antigen (c-kit proto-oncogene product) in normal human bone marrow. Blood 78:30-37.

AstraZeneca (2004). Faslodex, Fulvestrant Injection.

Atanassova N, McKinnell C, Turner KJ, Walker M, Fisher JS, Morley M, Millar MR, Groome NP, Sharpe RM (2000). Comparative effects of neonatal exposure of male rats to potent and weak (environmental) estrogens on spermatogenesis at puberty and the relationship to adult testis size and fertility: evidence for stimulatory effects of low estrogen levels. Endocrinology 141:3898-3907.

Awda BJ, Buhr MM (2010). Extracellular signal-regulated kinases (ERKs) pathway and reactive oxygen species regulate tyrosine phosphorylation in capacitating boar spermatozoa. Biol Reprod 83:750-758.

Basciani S, De LG, Dolci S, Brama M, Arizzi M, Mariani S, Rosano G, Spera G, Gnessi L (2008). Platelet-derived growth factor receptor beta-subtype regulates proliferation and migration of gonocytes. Endocrinology 149:6226-6235.

Basciani S, Mariani S, Arizzi M, Ulisse S, Rucci N, Jannini EA, Della RC, Manicone A, Carani C, Spera G, Gnessi L (2002). Expression of platelet-derived growth factor-A (PDGF-A), PDGF-B, and PDGF receptor-alpha and -beta during human testicular development and disease. J Clin Endocrinol Metab 87:2310-2319.

Basciani S, Mariani S, Spera G, Gnessi L (2010). Role of Platelet-Derived Growth Factors in the Testis. Endocr Rev.

Berensztein EB, Baquedano MS, Gonzalez CR, Saraco NI, Rodriguez J, Ponzio R, Rivarola MA, Belgorosky A (2006). Expression of aromatase, estrogen receptor alpha and beta, androgen receptor, and cytochrome P-450scc in the human early prepubertal testis. Pediatr Res 60:740-744.

Beumer TL, Roepers-Gajadien HL, Gademan IS, Kal HB, De Rooij DG (2000). Involvement of the D-type cyclins in germ cell proliferation and differentiation in the mouse. Biol Reprod 63:1893-1898.

100

Boulogne B, Habert R, Levacher C (2003). Regulation of the proliferation of cocultured gonocytes and Sertoli cells by retinoids, triiodothyronine, and intracellular signaling factors: differences between fetal and neonatal cells. Mol Reprod Dev 65:194-203.

Bouskine A, Nebout M, Mograbi B, Brucker-Davis F, Roger C, Fenichel P (2008). Estrogens promote human testicular germ cell cancer through a membrane-mediated activation of extracellular regulated kinase and protein kinase A. Endocrinology 149:565-573.

Brennan J, Tilmann C, Capel B (2003). Pdgfr-alpha mediates testis cord organization and fetal Leydig cell development in the XY gonad. Genes Dev 17:800-810.

Brinster RL, Zimmermann JW (1994). Spermatogenesis following male germ-cell transplantation. Proc Natl Acad Sci U S A 91:11298-11302.

Carreau S, Delalande C, Silandre D, Bourguiba S, Lambard S (2006). Aromatase and estrogen receptors in male reproduction. Mol Cell Endocrinol 246:65-68.

Carreau S, Lambard S, Delalande C, is-Galeraud I, Bilinska B, Bourguiba S (2003). Aromatase expression and role of estrogens in male gonad : a review. Reprod Biol Endocrinol 1:35.

Cheskis BJ, Greger J, Cooch N, McNally C, Mclarney S, Lam HS, Rutledge S, Mekonnen B, Hauze D, Nagpal S, Freedman LP (2008). MNAR plays an important role in ERa activation of Src/MAPK and PI3K/Akt signaling pathways. Steroids 73:901-905.

Choudhury GG, Karamitsos C, Hernandez J, Gentilini A, Bardgette J, Abboud HE (1997). PI-3-kinase and MAPK regulate mesangial cell proliferation and migration in response to PDGF. Am J Physiol 273:F931-F938.

Chuma S, Kanatsu-Shinohara M, Inoue K, Ogonuki N, Miki H, Toyokuni S, Hosokawa M, Nakatsuji N, Ogura A, Shinohara T (2005). Spermatogenesis from epiblast and primordial germ cells following transplantation into postnatal mouse testis. Development 132:117-122.

Collins P, Webb C (1999). Estrogen hits the surface. Nat Med 5:1130-1131.

Couse JF, Hewitt SC, Bunch DO, Sar M, Walker VR, Davis BJ, Korach KS (1999). Postnatal sex reversal of the ovaries in mice lacking estrogen receptors alpha and beta. Science 286:2328-2331.

Culig Z, Hobisch A, Cronauer MV, Radmayr C, Trapman J, Hittmair A, Bartsch G, Klocker H (1994). Androgen receptor activation in prostatic tumor cell lines by insulin-like growth factor-I, keratinocyte growth factor, and epidermal growth factor. Cancer Res 54:5474-5478.

Culty M (2009). Gonocytes, the forgotten cells of the germ cell lineage. Birth Defects Res C Embryo Today 87:1-26.

Culty M, Thuillier R, Li W, Wang Y, Martinez-Arguelles DB, Benjamin CG, Triantafilou KM, Zirkin BR, Papadopoulos V (2008). In utero exposure to di-(2-ethylhexyl) phthalate exerts both short-term and long-lasting suppressive effects on testosterone production in the rat. Biol Reprod 78:1018-1028.

101

Dai Y (2010). Platelet-derived growth factor receptor tyrosine kinase inhibitors: a review of the recent patent literature. Expert Opin Ther Pat 20:885-897.

Dai Z, Li Y, Quarles LD, Song T, Pan W, Zhou H, Xiao Z (2007). Resveratrol enhances proliferation and osteoblastic differentiation in human mesenchymal stem cells via ER-dependent ERK1/2 activation. Phytomedicine 14:806-814.

De Rooij DG (1998). Stem cells in the testis. Int J Exp Pathol 79:67-80.

De Rooij DG (2001). Proliferation and differentiation of spermatogonial stem cells. Reproduction 121:347-354.

de LE, O'Flaherty C (2008). Sperm activation: role of reactive oxygen species and kinases. Biochim Biophys Acta 1784:106-115.

Delbes G, Duquenne C, Szenker J, Taccoen J, Habert R, Levacher C (2007). Developmental changes in testicular sensitivity to estrogens throughout fetal and neonatal life. Toxicol Sci 99:234-243.

Delbes G, Levacher C, Habert R (2006). Estrogen effects on fetal and neonatal testicular development. Reproduction 132:527-538.

Delbes G, Levacher C, Pairault C, Racine C, Duquenne C, Krust A, Habert R (2004). Estrogen receptor beta-mediated inhibition of male germ cell line development in mice by endogenous estrogens during perinatal life. Endocrinology 145:3395-3403.

Dong L, Jelinsky SA, Finger JN, Johnston DS, Kopf GS, Sottas CM, Hardy MP, Ge RS (2007). Gene expression during development of fetal and adult Leydig cells. Ann N Y Acad Sci 1120:16-35.

Eddy EM, Washburn TF, Bunch DO, Goulding EH, Gladen BC, Lubahn DB, Korach KS (1996). Targeted disruption of the estrogen receptor gene in male mice causes alteration of spermatogenesis and infertility. Endocrinology 137:4796-4805.

Fox EM, Andrade J, Shupnik MA (2009). Novel actions of estrogen to promote proliferation: integration of cytoplasmic and nuclear pathways. Steroids 74:622-627.

Fu XD, Cui YH, Lin GP, Wang TH (2007). Non-genomic effects of 17beta-estradiol in activation of the ERK1/ERK2 pathway induces cell proliferation through upregulation of cyclin D1 expression in bovine artery endothelial cells. Gynecol Endocrinol 23:131-137.

Gaskell TL, Esnal A, Robinson LL, Anderson RA, Saunders PT (2004). Immunohistochemical profiling of germ cells within the human fetal testis: identification of three subpopulations. Biol Reprod 71:2012-2021.

Geffroy N, Guedin A, Dacquet C, Lefebvre P (2005). Cell cycle regulation of breast cancer cells through estrogen-induced activities of ERK and Akt protein kinases. Mol Cell Endocrinol 237:11-23.

Gnessi L, Emidi A, Jannini EA, Carosa E, Maroder M, Arizzi M, Ulisse S, Spera G (1995). Testicular development involves the spatiotemporal control of PDGFs and PDGF receptors gene expression and action. J Cell Biol 131:1105-1121.

102

Gould ML, Hurst PR, Nicholson HD (2007). The effects of oestrogen receptors alpha and beta on testicular cell number and steroidogenesis in mice. Reproduction 134:271-279.

Gray K, Eitzman B, Raszmann K, Steed T, Geboff A, McLachlan J, Bidwell M (1995). Coordinate regulation by diethylstilbestrol of the platelet-derived growth factor-A (PDGF-A) and -B chains and the PDGF receptor alpha- and beta-subunits in the mouse uterus and vagina: potential mediators of estrogen action. Endocrinology 136:2325-2340.

Gutendorf B, Westendorf J (2001). Comparison of an array of in vitro assays for the assessment of the estrogenic potential of natural and synthetic estrogens, phytoestrogens and xenoestrogens. Toxicology 166:79-89.

Habert R, Lejeune H, Saez JM (2001). Origin, differentiation and regulation of fetal and adult Leydig cells. Mol Cell Endocrinol 179:47-74.

He X, Andersson G, Lindgren U, Li Y (2010). Resveratrol prevents RANKL-induced osteoclast differentiation of murine osteoclast progenitor RAW 264.7 cells through inhibition of ROS production. Biochem Biophys Res Commun 401:356-362.

Heinrich MC, Corless CL, Duensing A, McGreevey L, Chen CJ, Joseph N, Singer S, Griffith DJ, Haley A, Town A, Demetri GD, Fletcher CD, Fletcher JA (2003). PDGFRA activating mutations in gastrointestinal stromal tumors. Science 299:708-710.

Heldring N, Pike A, Andersson S, Matthews J, Cheng G, Hartman J, Tujague M, Strom A, Treuter E, Warner M, Gustafsson JA (2007). Estrogen receptors: how do they signal and what are their targets. Physiol Rev 87:905-931.

Hess RA, Carnes K (2004). The Role of Estrogen in Testis and the Male Reproductive Tract: a Review and Species Comparison. Animal Reproduction 1:5-30.

Hilscher W (1991). The genetic control and germ cell kinetics of the female and male germ line in mammals including man. Hum Reprod 6:1416-1425.

Hirahara Y, Matsuda K, Gao W, Arvanitis DN, Kawata M, Boggs JM (2009). The localization and non-genomic function of the membrane-associated estrogen receptor in oligodendrocytes. Glia 57:153-165.

Hoei-Hansen CE, Olesen IA, Jorgensen N, Carlsen E, Holm M, Almstrup K, Leffers H, Rajpert-De ME (2007). Current approaches for detection of carcinoma in situ testis. Int J Androl 30:398-404.

Holmes L, Jr., Escalante C, Garrison O, Foldi BX, Ogungbade GO, Essien EJ, Ward D (2008). Testicular cancer incidence trends in the USA (1975-2004): plateau or shifting racial paradigm? Public Health 122:862-872.

Horwich A, Shipley J, Huddart R (2006). Testicular germ-cell cancer. Lancet 367:754-765.

Huckins C (1971). The spermatogonial stem cell population in adult rats. I. Their morphology, proliferation and maturation. Anat Rec 169:533-557.

103

Ignar-Trowbridge DM, Nelson KG, Bidwell MC, Curtis SW, Washburn TF, McLachlan JA, Korach KS (1992). Coupling of dual signaling pathways: epidermal growth factor action involves the estrogen receptor. Proc Natl Acad Sci U S A 89:4658-4662.

Ignar-Trowbridge DM, Pimentel M, Teng CT, Korach KS, McLachlan JA (1995). Cross talk between peptide growth factor and estrogen receptor signaling systems. Environ Health Perspect 103 Suppl 7:35-38.

Iguchi T, Watanabe H, Katsu Y (2001). Developmental effects of estrogenic agents on mice, fish, and frogs: a mini-review. Horm Behav 40:248-251.

Jahnukainen K, Chrysis D, Hou M, Parvinen M, Eksborg S, Soder O (2004). Increased apoptosis occurring during the first wave of spermatogenesis is stage-specific and primarily affects midpachytene spermatocytes in the rat testis. Biol Reprod 70:290-296.

Jensen EV (2004). From Chemical Warfare to Breast Cancer Management. p. 1018-1021.

Joensen UN, Jorgensen N, Skakkebaek NE (2007). Testicular dysgenesis syndrome and carcinoma in situ of the testes. Nat Clin Pract Urol 4:402-403.

Johnston DS, Russell LD, Friel PJ, Griswold MD (2001). Murine germ cells do not require functional androgen receptors to complete spermatogenesis following spermatogonial stem cell transplantation. Endocrinology 142:2405-2408.

Kaarbo M, Klokk TI, Saatcioglu F (2007). Androgen signaling and its interactions with other signaling pathways in prostate cancer. Bioessays 29:1227-1238.

Kim EK, Choi EJ (2010). Pathological roles of MAPK signaling pathways in human diseases. Biochim Biophys Acta 1802:396-405.

Koehler KF, Helguero LA, Haldosen LA, Warner M, Gustafsson JA (2005). Reflections on the discovery and significance of estrogen receptor beta. Endocr Rev 26:465-478.

Kolch W (2000). Meaningful relationships: the regulation of the Ras/Raf/MEK/ERK pathway by protein interactions. Biochem J 351 Pt 2:289-305.

Krege JH, Hodgin JB, Couse JF, Enmark E, Warner M, Mahler JF, Sar M, Korach KS, Gustafsson JA, Smithies O (1998). Generation and reproductive phenotypes of mice lacking estrogen receptor beta. Proc Natl Acad Sci U S A 95:15677-15682.

Kupisiewicz K, Boissy P, Abdallah BM, Hansen FD, Erben RG, Savouret JF, Soe K, Andersen TL, Plesner T, Delaisse JM (2010). Potential of resveratrol analogues as antagonists of osteoclasts and promoters of osteoblasts. Calcif Tissue Int 87:437-449.

La Sala G, Farini D, De FM (2010). Rapid estrogen signalling in mouse primordial germ cells. Exp Cell Res 316:1716-1727.

Lambrot R, Coffigny H, Pairault C, Donnadieu AC, Frydman R, Habert R, Rouiller-Fabre V (2006). Use of organ culture to study the human fetal testis development: effect of retinoic acid. J Clin Endocrinol Metab 91:2696-2703.

104

Lau WS, Chen WF, Chan RY, Guo DA, Wong MS (2009). Mitogen-activated protein kinase (MAPK) pathway mediates the oestrogen-like activities of ginsenoside Rg1 in human breast cancer (MCF-7) cells. Br J Pharmacol 156:1136-1146.

Lawson KA, Dunn NR, Roelen BA, Zeinstra LM, Davis AM, Wright CV, Korving JP, Hogan BL (1999). Bmp4 is required for the generation of primordial germ cells in the mouse embryo. Genes Dev 13:424-436.

Lazari MF, Lucas TF, Yasuhara F, Gomes GR, Siu ER, Royer C, Fernandes SA, Porto CS (2009). Estrogen receptors and function in the male reproductive system. Arq Bras Endocrinol Metabol 53:923-933.

Li H, Papadopoulos V, Vidic B, Dym M, Culty M (1997). Regulation of rat testis gonocyte proliferation by platelet-derived growth factor and estradiol: identification of signaling mechanisms involved. Endocrinology 138:1289-1298.

Livera G, Delbes G, Pairault C, Rouiller-Fabre V, Habert R (2006). Organotypic culture, a powerful model for studying rat and mouse fetal testis development. Cell Tissue Res 324:507-521.

Livera G, Rouiller-Fabre V, Durand P, Habert R (2000). Multiple effects of retinoids on the development of Sertoli, germ, and Leydig cells of fetal and neonatal rat testis in culture. Biol Reprod 62:1303-1314.

Lu X, Qu CK, Shi ZQ, Feng GS (1998). Downregulation of platelet-derived growth factor receptor-beta in Shp-2 mutant fibroblast cell lines. Oncogene 17:441-448.

Luconi M, Forti G, Baldi E (2002). Genomic and nongenomic effects of estrogens: molecular mechanisms of action and clinical implications for male reproduction. J Steroid Biochem Mol Biol 80:369-381.

MacCalman CD, Getsios S, Farookhi R, Blaschuk OW (1997). Estrogens potentiate the stimulatory effects of follicle-stimulating hormone on N-cadherin messenger ribonucleic acid levels in cultured mouse Sertoli cells. Endocrinology 138:41-48.

Maggiolini M, Picard D (2010). The unfolding stories of GPR30, a new membrane-bound estrogen receptor. J Endocrinol 204:105-114.

Maher B (2006). The Discovery of Estrogen Receptor Beta. p. 92.

Mariani S, Basciani S, Arizzi M, Spera G, Gnessi L (2002). PDGF and the testis. Trends Endocrinol Metab 13:11-17.

Martelli AM, Evangelisti C, Chiarini F, McCubrey JA (2010). The phosphatidylinositol 3-kinase/Akt/mTOR signaling network as a therapeutic target in acute myelogenous leukemia patients. Oncotarget 1:89-103.

McGuinness MP, Orth JM (1992). Reinitiation of gonocyte mitosis and movement of gonocytes to the basement membrane in testes of newborn rats in vivo and in vitro. Anat Rec 233:527-537.

McIntyre A, Summersgill B, Grygalewicz B, Gillis AJ, Stoop J, van Gurp RJ, Dennis N, Fisher C, Huddart R, Cooper C, Clark J, Oosterhuis JW, Looijenga LH, Shipley J (2005). Amplification and overexpression of the KIT gene is associated with progression in the seminoma subtype of testicular germ cell tumors of adolescents and adults. Cancer Res 65:8085-8089.

105

McLaren A (1999). Signaling for germ cells. Genes Dev 13:373-376.

Mercier I, Mader S, Calderone A (2003). Tamoxifen and ICI 182,780 negatively influenced cardiac cell growth via an estrogen receptor-independent mechanism. Cardiovasc Res 59:883-892.

Merlet J, Racine C, Moreau E, Moreno SG, Habert R (2007). Male fetal germ cells are targets for androgens that physiologically inhibit their proliferation. Proc Natl Acad Sci U S A 104:3615-3620.

Migliaccio A, Castoria G, Di DM, de FA, Bilancio A, Lombardi M, Barone MV, Ametrano D, Zannini MS, Abbondanza C, Auricchio F (2000). Steroid-induced androgen receptor-oestradiol receptor beta-Src complex triggers prostate cancer cell proliferation. EMBO J 19:5406-5417.

Mitsunaga K, Araki K, Mizusaki H, Morohashi K, Haruna K, Nakagata N, Giguere V, Yamamura K, Abe K (2004). Loss of PGC-specific expression of the orphan nuclear receptor ERR-beta results in reduction of germ cell number in mouse embryos. Mech Dev 121:237-246.

Moreno SG, Dutrillaux B, Coffigny H (2001). Status of p53, p21, mdm2, pRb proteins, and DNA methylation in gonocytes of control and gamma-irradiated rats during testicular development. Biol Reprod 64:1422-1431.

Mosselman S, Claesson-Welsh L, Kamphuis JS, van Zoelen EJ (1994). Developmentally regulated expression of two novel platelet-derived growth factor alpha-receptor transcripts in human teratocarcinoma cells. Cancer Res 54:220-225.

Mosselman S, Looijenga LH, Gillis AJ, van Rooijen MA, Kraft HJ, van Zoelen EJ, Oosterhuis JW (1996). Aberrant platelet-derived growth factor alpha-receptor transcript as a diagnostic marker for early human germ cell tumors of the adult testis. Proc Natl Acad Sci U S A 93:2884-2888.

Nadal A, Ropero AB, Laribi O, Maillet M, Fuentes E, Soria B (2000). Nongenomic actions of estrogens and xenoestrogens by binding at a plasma membrane receptor unrelated to estrogen receptor alpha and estrogen receptor beta. Proc Natl Acad Sci U S A 97:11603-11608.

Nagano R, Tabata S, Nakanishi Y, Ohsako S, Kurohmaru M, Hayashi Y (2000). Reproliferation and relocation of mouse male germ cells (gonocytes) during prespermatogenesis. Anat Rec 258:210-220.

Nikov GN, Eshete M, Rajnarayanan RV, Alworth WL (2001). Interactions of synthetic estrogens with human estrogen receptors. J Endocrinol 170:137-145.

Nitta H, Bunick D, Hess RA, Janulis L, Newton SC, Millette CF, Osawa Y, Shizuta Y, Toda K, Bahr JM (1993). Germ cells of the mouse testis express P450 aromatase. Endocrinology 132:1396-1401.

Nurmio M, Kallio J, Toppari J, Jahnukainen K (2008). Adult reproductive functions after early postnatal inhibition by imatinib of the two receptor tyrosine kinases, c-kit and PDGFR, in the rat testis. Reprod Toxicol 25:442-446.

Nurmio M, Toppari J, Zaman F, Andersson AM, Paranko J, Soder O, Jahnukainen K (2007). Inhibition of tyrosine kinases PDGFR and C-Kit by imatinib mesylate interferes with postnatal testicular development in the rat. Int J Androl 30:366-376.

106

O'Flaherty C, de LE, Gagnon C (2006). Positive role of reactive oxygen species in mammalian sperm capacitation: triggering and modulation of phosphorylation events. Free Radic Biol Med 41:528-540.

Oakberg EF (1971). Spermatogonial stem-cell renewal in the mouse. Anat Rec 169:515-531.

Ohbo K, Yoshida S, Ohmura M, Ohneda O, Ogawa T, Tsuchiya H, Kuwana T, Kehler J, Abe K, Scholer HR, Suda T (2003). Identification and characterization of stem cells in prepubertal spermatogenesis in mice small star, filled. Dev Biol 258:209-225.

Ohmura M, Yoshida S, Ide Y, Nagamatsu G, Suda T, Ohbo K (2004). Spatial analysis of germ stem cell development in Oct-4/EGFP transgenic mice. Arch Histol Cytol 67:285-296.

Ohta H, Wakayama T, Nishimune Y (2004). Commitment of fetal male germ cells to spermatogonial stem cells during mouse embryonic development. Biol Reprod 70:1286-1291.

Olaso R, Habert R (2000). Genetic and cellular analysis of male germ cell development. J Androl 21:497-511.

Orth JM, Boehm R (1990). Functional coupling of neonatal rat Sertoli cells and gonocytes in coculture. Endocrinology 127:2812-2820.

Osborne CK, Shou J, Massarweh S, Schiff R (2005). Crosstalk between estrogen receptor and growth factor receptor pathways as a cause for endocrine therapy resistance in breast cancer. Clin Cancer Res 11:865s-870s.

Palumbo C, van RK, Gillis AJ, van Gurp RH, de MH, Oosterhuis JW, van Zoelen EJ, Looijenga LH (2002). Expression of the PDGF alpha-receptor 1.5 kb transcript, OCT-4, and c-KIT in human normal and malignant tissues. Implications for the early diagnosis of testicular germ cell tumours and for our understanding of regulatory mechanisms. J Pathol 196:467-477.

Park JH, Lee MY, Han HJ (2009). A potential role for caveolin-1 in estradiol-17beta-induced proliferation of mouse embryonic stem cells: involvement of Src, PI3K/Akt, and MAPKs pathways. Int J Biochem Cell Biol 41:659-665.

Pettersson K, Gustafsson JA (2001). Role of estrogen receptor beta in estrogen action. Annu Rev Physiol 63:165-192.

Puglianiello A, Campagnolo L, Farini D, Cipollone D, Russo MA, Siracusa G (2004). Expression and role of PDGF-BB and PDGFR-beta during testis morphogenesis in the mouse embryo. J Cell Sci 117:1151-1160.

Rajpert-De Meyts E., Hoei-Hansen CE (2007). From gonocytes to testicular cancer: the role of impaired gonadal development. Ann N Y Acad Sci 1120:168-180.

Ricci G, Catizone A, Galdieri M (2004). Embryonic mouse testis development: role of platelet derived growth factor (PDGF-BB). J Cell Physiol 200:458-467.

Robinson MJ, Cobb MH (1997). Mitogen-activated protein kinase pathways. Curr Opin Cell Biol 9:180-186.

107

Roosen-Runge EC, Leik J (1968). Gonocyte degeneration in the postnatal male rat. Am J Anat 122:275-299.

Rouiller-Fabre V, Levacher C, Pairault C, Racine C, Moreau E, Olaso R, Livera G, Migrenne S, Delbes G, Habert R (2003). Development of the foetal and neonatal testis. Andrologia 35:79-83.

Rumora L, Grubisic TZ (2009). A journey through mitogen-activated protein kinase and ochratoxin A interactions. Arh Hig Rada Toksikol 60:449-456.

Russell LD, Ettlin R, Sinha Hikim A, Clegg E (1990) Histological and Histopathological Evaluation of the Testis.Cache River Press.

Schaeffer HJ, Weber MJ (1999). Mitogen-activated protein kinases: specific messages from ubiquitous messengers. Mol Cell Biol 19:2435-2444.

Schlatt S (2010) Schlatt Laboratory.

Schmahl J, Rizzolo K, Soriano P (2008). The PDGF signaling pathway controls multiple steroid-producing lineages. Genes Dev 22:3255-3267.

Segars JH, Driggers PH (2002). Estrogen action and cytoplasmic signaling cascades. Part I: membrane-associated signaling complexes. Trends Endocrinol Metab 13:349-354.

Sehmisch S, Hammer F, Christoffel J, Seidlova-Wuttke D, Tezval M, Wuttke W, Stuermer KM, Stuermer EK (2008). Comparison of the phytohormones genistein, resveratrol and 8-prenylnaringenin as agents for preventing osteoporosis. Planta Med 74:794-801.

Sharpe RM, Skakkebaek NE (2008). Testicular dysgenesis syndrome: mechanistic insights and potential new downstream effects. Fertil Steril 89:e33-e38.

Shaul YD, Seger R (2007). The MEK/ERK cascade: from signaling specificity to diverse functions. Biochim Biophys Acta 1773:1213-1226.

Simoncini T, Scorticati C, Mannella P, Fadiel A, Giretti MS, Fu XD, Baldacci C, Garibaldi S, Caruso A, Fornari L, Naftolin F, Genazzani AR (2006). Estrogen receptor alpha interacts with Galpha13 to drive actin remodeling and endothelial cell migration via the RhoA/Rho kinase/moesin pathway. Mol Endocrinol 20:1756-1771.

Soares R, Guerreiro S, Botelho M (2007). Elucidating progesterone effects in breast cancer: cross talk with PDGF signaling pathway in smooth muscle cell. J Cell Biochem 100:174-183.

Sonne SB, Kristensen DM, Novotny GW, Olesen IA, Nielsen JE, Skakkebaek NE, Rajpert-De ME, Leffers H (2008). Testicular dysgenesis syndrome and the origin of carcinoma in situ testis. Int J Androl 31:275-287.

Strauss J, Barbieri R (2004) Reproductive Endocrinology.Elsevier Saunders.

Sutton KA (2000). Molecular mechanisms involved in the differentiation of spermatogenic stem cells. Rev Reprod 5:93-98.

108

Thuillier R, Manku G, Wang Y, Culty M (2009). Changes in MAPK pathway in neonatal and adult testis following fetal estrogen exposure and effects on rat testicular cells. Microsc Res Tech 72:773-786.

Thuillier R, Mazer M, Manku G, Boisvert A, Wang Y, Culty M (2010). Interdependence of platelet-derived growth factor and estrogen-signaling pathways in inducing neonatal rat testicular gonocytes proliferation. Biol Reprod 82:825-836.

Thuillier R, Wang Y, Culty M (2003). Prenatal exposure to estrogenic compounds alters the expression pattern of platelet-derived growth factor receptors alpha and beta in neonatal rat testis: identification of gonocytes as targets of estrogen exposure. Biol Reprod 68:867-880.

Tres LL, Kierszenbaum AL (2005). The ADAM-integrin-tetraspanin complex in fetal and postnatal testicular cords. Birth Defects Res C Embryo Today 75:130-141.

van Dissel-Emiliani FM, De Rooij DG, Meistrich ML (1989). Isolation of rat gonocytes by velocity sedimentation at unit gravity. J Reprod Fertil 86:759-766.

Vigueras-Villasenor RM, Moreno-Mendoza NA, Reyes-Torres G, Molina-Ortiz D, Leon MC, Rojas-Castaneda JC (2006). The effect of estrogen on testicular gonocyte maturation. Reprod Toxicol 22:513-520.

Vornberger W, Prins G, Musto NA, Suarez-Quian CA (1994). Androgen receptor distribution in rat testis: new implications for androgen regulation of spermatogenesis. Endocrinology 134:2307-2316.

Vu TH, Martin GR, Lee P, Mark D, Wang A, Williams LT (1989). Developmentally regulated use of alternative promoters creates a novel platelet-derived growth factor receptor transcript in mouse teratocarcinoma and embryonic stem cells. Mol Cell Biol 9:4563-4567.

Wahlgren A, Svechnikov K, Strand ML, Jahnukainen K, Parvinen M, Gustafsson JA, Soder O (2008). Estrogen receptor beta selective ligand 5alpha-Androstane-3beta, 17beta-diol stimulates spermatogonial deoxyribonucleic acid synthesis in rat seminiferous epithelium in vitro. Endocrinology 149:2917-2922.

Wang Y, Culty M (2007). Identification and distribution of a novel platelet-derived growth factor receptor beta variant: effect of retinoic acid and involvement in cell differentiation. Endocrinology 148:2233-2250.

Wang Y, Thuillier R, Culty M (2004). Prenatal estrogen exposure differentially affects estrogen receptor-associated proteins in rat testis gonocytes. Biol Reprod 71:1652-1664.

Waterloo (2007). Biology of the Female Reproductive system.

Wong CW, McNally C, Nickbarg E, Komm BS, Cheskis BJ (2002). Estrogen receptor-interacting protein that modulates its nongenomic activity-crosstalk with Src/Erk phosphorylation cascade. Proc Natl Acad Sci U S A 99:14783-14788.

Ying Y, Liu XM, Marble A, Lawson KA, Zhao GQ (2000). Requirement of Bmp8b for the generation of primordial germ cells in the mouse. Mol Endocrinol 14:1053-1063.

109

Yoshida S, Sukeno M, Nabeshima Y (2007). A vasculature-associated niche for undifferentiated spermatogonia in the mouse testis. Science 317:1722-1726.

Yoshida S, Sukeno M, Nakagawa T, Ohbo K, Nagamatsu G, Suda T, Nabeshima Y (2006). The first round of mouse spermatogenesis is a distinctive program that lacks the self-renewing spermatogonia stage. Development 133:1495-1505.

Zhang L, Li X, Zhao L, Zhang L, Zhang G, Wang J, Wei L (2009). Nongenomic effect of estrogen on the MAPK signaling pathway and calcium influx in endometrial carcinoma cells. J Cell Biochem 106:553-562.

Zhao L, O'Neill K, Brinton RD (2006). Estrogenic agonist activity of ICI 182,780 (Faslodex) in hippocampal neurons: implications for basic science understanding of estrogen signaling and development of estrogen modulators with a dual therapeutic profile. J Pharmacol Exp Ther 319:1124-1132.

Zhou Q, Li Y, Nie R, Friel P, Mitchell D, Evanoff RM, Pouchnik D, Banasik B, McCarrey JR, Small C, Griswold MD (2008a). Expression of stimulated by retinoic acid gene 8 (Stra8) and maturation of murine gonocytes and spermatogonia induced by retinoic acid in vitro. Biol Reprod 78:537-545.

Zhou Q, Nie R, Li Y, Friel P, Mitchell D, Hess RA, Small C, Griswold MD (2008b). Expression of stimulated by retinoic acid gene 8 (Stra8) in spermatogenic cells induced by retinoic acid: an in vivo study in vitamin A-sufficient postnatal murine testes. Biol Reprod 79:35-42.

110