Mechanisms of tumour suppression and control › record › 5337 › files ›...

154
UNIVERSITÉ DE GENÈVE Département de biologie moléculaire FACULTÉ DES SCIENCES Professeur U. Schibler Département de réhabilitation et gériatrie FACULTÉ DE MÉDECINE Professeur K.-H. Krause Docteur I. Irminger-Finger Mechanisms of Tumour Suppression and Control of Genomic Integrity by BARD1 THÈSE Présenté à la faculté des sciences de l’Université de Genève pour obtenir le grade de Docteur ès sciences, mention biologique par Charles Edward Jefford de Troinex (GE) Thèse N° 3651 Genève, 2005

Transcript of Mechanisms of tumour suppression and control › record › 5337 › files ›...

Page 1: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

UNIVERSITÉ DE GENÈVE Département de biologie moléculaire FACULTÉ DES SCIENCES Professeur U. Schibler Département de réhabilitation et gériatrie FACULTÉ DE MÉDECINE Professeur K.-H. Krause Docteur I. Irminger-Finger

Mechanisms of Tumour Suppression and Control of Genomic Integrity by BARD1

THÈSE Présenté à la faculté des sciences de l’Université de Genève

pour obtenir le grade de Docteur ès sciences, mention biologique

par

Charles Edward Jefford

de

Troinex (GE)

Thèse N° 3651 Genève, 2005

Page 2: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

UNIVERSITÉ DE GENÈVE Département de biologie moléculaire FACULTÉ DES SCIENCES Professeur U. Schibler Département de réhabilitation et gériatrie FACULTÉ DE MÉDECINE Professeur K.-H. Krause Docteur I. Irminger-Finger

Mechanisms of Tumour Suppression and Control of Genomic Integrity by BARD1

THÈSE Présenté à la faculté des sciences de l’Université de Genève

pour obtenir le grade de Docteur ès sciences, mention biologique

par

Charles Edward Jefford

de

Troinex (GE)

Thèse N° 3651 Genève, 2005

Page 3: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

FACULTÉDESSCIENCES

..

DE GENEVE

Doctorat ès sciencesmention biologique

Thèse de Monsieur Charles Edward JEFFORD

intitulée:

"Mechanisms of Tumour Suppression and Control ofGenomic Integrity by BARD1 Il

La Faculté des sciences, sur le préavis de Monsieur K.-H. KRAUSE,professeur ordinaire et

directeur de thèse (Département de gériatrie - laboratoire de biologie du vieillissement),de Madame 1. IRMINGER-FINGER,docteur (Département de gériatrie - laboratoire de

biologie du vieillissement), et U. SCHIBLER,professeur ordinaire (Département de biologie

moléculaire) co-directeurs de thèse, et M. AGUET, professeur (Institut Suissede recherche

expérimentale surle cancer - Epalinges,Suisse),autorise l'impressionde la présente thèse,sans exprimer d'opinion sur les propositions qui y sont énoncées.

Genève, le 18 août 2005

(1 ~ Y'" \Vl V'Thèse - 3651 -

Le Doyen, Pierre SPIERER

N.B.- La thèse doit porter la déclaration précédente et remplir les conditions énuméréesdans les "Informations relatives aux thèses de doctorat à l'Université de Genève".

Nombre d'exemplaires à livrer par colis séparé à la Faculté: - 7 -

Page 4: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

ii

This thesis is dedicated to the ones I love: to my family and especially to my

parents who unstintingly supported me throughout my university years.

Page 5: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

iii

Acknowledgements

I would like to thank my thesis supervisor and director, Dr Irmgard Irminger-Finger, for

being such an excellent guide through this thesis and for her continuous enthusiasm;

Professor Karl-Heinz Krause for being my thesis director and for welcoming me in the

Laboratory of Aging Biology. I would like to thank Professors Ueli Schibler and Michel

Aguet for accepting to be my thesis jury; Dr Pedro Herrera and Dr Michael Pepper for

accepting to be my “parrains de thèse”.

I would like to address my appreciation to Dr Anis Feki, Dr Esther Bettiol, Dr Laetitia

Cartier-Fioraso and Dr Stephan Ryser for their friendship, advice and support; to

Aurélie Caillon, Laura Ortolan-Trigui, Chantal Genet, Terese Laforge and Olivier

Plastre for their kindness, support and excellent technical assistance; Dr Jacques

Proust and Dr Sophie Clement-Leboube for helpful comments on the manuscript; and

Hélène Gfeller-Tillmann for administrative help. I would like to thank Dr Sarantis

Gagos for providing his expertise on cytogenetic analysis, Dr Jean Harb on electronic

microscopy, Dr Serge Arnaudeau on confocal microscopy, Dr Bénédicte Delaval and

Dr Daniel Birnbaum for their precious collaboration.

I would like to give credit to all previous and present members of the Laboratory of

Aging Biology and collaborators who helped me in various ways: Dr Marisa Jaconi, Dr

Michel Dubois-Dauphin, Dr Lena Serrander, Dr Igor Bondarev, Dr Gaetan Gavazzi, Dr

Philippe Berardi, Dr Jian Li, Dr Botond Banfi, Dr Mathieu Hauwel, Dr Michela

Schaeppi, Dr Karen Bedard, Dr Anne-Thérèse Vlastos, Nicolas Molnarfi, Stéphanie

Julien, Dr David Suter, Michael Stouffs, Fabrice Calabria, Mamadou Hadi-Sow, Dr

Jian Yu Wu, Christine Deffert, Diderik Tirefort, Paula Borel, Françoise Piotton and

Silvia Sorce.

Page 6: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

iv

Furthermore, I would like to acknowledge my former mentors Dr Richard Isbrucker, Dr

Ross Longley and Dr Shirley Pomponi at the Harbor Branch Oceanographic

Institution in Florida who were instrumental in rekindling my interest in biological

sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided the

opportunity to work at HBOI.

Finally, I would like to express my gratitude to my father, Professor Charles William

Jefford, for his unstinting support and encouragement, and for kindly reading the

manuscript. I would also like to acknowledge my mother, sisters, nieces and brothers-

in-law for their constant support: Susan, Sarah, Alex, Kasmira, Nina, Nieve, Mikael

and Hossein without whom this doctorate would certainly not have been possible.

Page 7: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

v

TABLE OF CONTENTS page

Avant propos. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 Organisation de la thèse. . . . . . . . . . . . . . . . . . . . . . . . . . 1

Résumé de la thèse . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

Introduction sur BARD1 . . . . . . . . . . . . . . . . . . . . . . . . . . 2

Structure de BARD1. . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

Les modèles génétiques . . . . . . . . . . . . . . . . . . . . . . . . . 3

L’activité ubiquitine ligase de BARD1. . . . . . . . . . . . . . . . . . . 3

Les fonctions de BARD1 dépendantes de BRCA1. . . . . . . . . . . . 4

Les fonctions de BARD1 indépendantes de BRCA1. . . . . . . . . . . 5

Les objectifs de la thèse. . . . . . . . . . . . . . . . . . . . . . . . . . 5

Résultats et discussion. . . . . . . . . . . . . . . . . . . . . . . . . . . 6

Conclusion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

Preface. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

The purpose of cancer. . . . . . . . . . . . . . . . . . . . . . . . . . . 14

Review on BARD1 and its diverse functions. . . . . . . . . . . . . . . . . 16

I) Breast cancer and the BRCA tumour suppressors. . . . . . . . . . . 16

BRCA1 and BRCA2 mutations in cancer. . . . . . . . . . . . . . 16

Genetic models for BRCA1 and BRCA2. . . . . . . . . . . . . . 18

Overview BRCA1 and BRCA2 functions. . . . . . . . . . . . . . 18

II) BARD1 protein structure . . . . . . . . . . . . . . . . . . . . . . . . 19

BARD1 discovery and phylogenetic analysis. . . . . . . . . . . . 19

BARD1 orthologues. . . . . . . . . . . . . . . . . . . . . . . . . 21

RING finger domain. . . . . . . . . . . . . . . . . . . . . . . . . 23

Ankyrin repeats. . . . . . . . . . . . . . . . . . . . . . . . . . . 25

BRCT domains. . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

III) Ubiquitin ligase functions of BRCA1/BARD1. . . . . . . . . . . . . . 26

RING finger domain proteins are ubiquitin ligases. . . . . . . . . 26

Ubiquitination and de-ubiquitination. . . . . . . . . . . . . . . . . 26

BARD1/BRCA1 heterodimer is an E3 ubiquitin ligase. . . . . . . 27

Ubiquitin targets of the BARD1/BRCA1 heterodimer. . . . . . . . 28

IV) BARD1 functions and its associated proteins. . . . . . . . . . . . . . 31

Page 8: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

vi

Knockout experiments and genomic instability phenotype. . . . . 32

BARD1 in DNA-damage repair. . . . . . . . . . . . . . . . . . . 33

Cellular localisation of BARD1 and BRCA1 during DNA repair. . 33

BARD1 in transcriptional regulation and chromatin remodelling. . 34

BARD1 in mRNA processing. . . . . . . . . . . . . . . . . . . . 36

BRCA1-independent pro-apoptotic functions of BARD1 . . . . . . 37

Mapping of pro-apoptotic domain of BARD1. . . . . . . . . . . . . 39

BRCA1 and BARD1 functions in cell cycle. . . . . . . . . . . . . 40

V) BARD1 mutations and expression in cancer. . . . . . . . . . . . . . 41

BARD1 tumour suppressor mutations in breast and ovarian cancers 41

BARD1 expression in normal and tumour tissues. . . . . . . . . 43

VI) Conclusion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

Results. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

Published results (1 to 4) and paper submitted for publication (5) are preceded

by a one page summary and explain my contribution in each project.

Supplemental results (6 and 7) are summarized and accompanied by figures.

Publications:

1. Irminger-Finger I, Leung WC, Li J, Dubois-Dauphin M,Harb J, Feki A,

Jefford CE, Soriano JV, Jaconi M, Montessano R, Krause KH. Identification

of BARD1 as mediator between proapoptotic stress and p53-dependent

apoptosis. Mol Cell. 2001 Dec;8(6):1255-66. . . . . . . . . . . . . . . . . . 48

2. Jefford CE, Feki A, Harb J, Krause KH, Irminger-Finger I. Nuclear-

cytoplasmic translocation of BARD1 is linked to its apoptotic activity.

Oncogene. 2004 Apr 29;23(20):3509-20. . . . . . . . . . . . . . . . . . . 60

3. Feki A, Jefford CE, Durand P, Harb J, Lucas H, Krause KH, Irminger-Finger

I. BARD1 expression during spermatogenesis is associated with apoptosis

and hormonally regulated. Biol Reprod. 2004 Nov;71(5):1614-24. . . . . . . 73

4. Feki A, Jefford CE, Berardi P, Wu JY, Cartier L, Krause KH, Irminger-

Finger I. BARD1 induces apoptosis by catalyzing phosphorylation of p53 by

DNA-damage response kinase. Oncogene. 2005 Mar 14 (in press). . . . . . 85

Page 9: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

vii

Submitted for publication:

5. Jefford CE, Delaval B, Ryser S, Birnbaum D, Irminger-Finger I.

BARD1 interacts with BRCA2 to regulate cytokinesis and maintain genomic

stability. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97

Supplemental results and ongoing projects:

6. Role of BARD1 in telomere maintenance and crisis. . . . . . . . . . . . . . 124

7. BARD1 promoter analysis. . . . . . . . . . . . . . . . . . . . . . . . . . . 130

Conclusion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136

References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137

Page 10: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

1

Avant propos Organisation de la thèse

Quelques mots sont nécessaires pour décrire le plan de la thèse. La thèse est

divisée en plusieurs parties. Le premier chapitre est en français, résume toute la

thèse et constitue aussi l’introduction de thèse. Il contient un résumé de la revue sur

BARD1, les objectifs et le plan de la thèse, un résumé précis des résultats et

discussions, divisés en 7 sous-chapitres, ainsi qu’une petite conclusion. Le reste de

la thèse est en anglais à cause des publications. Le deuxième chapitre est une

préface philosophique sur le rôle des tumeurs dans la population. Le troisième

chapitre est une revue complète sur BARD1 qui sera prochainement soumis pour

une publication dans un journal et qui a déjà servi en partie comme chapitre pour un

livre. Le quatrième chapitre regroupe tous les résultats et est divisé en 7 parties

distinctes. Ces parties sont reliées entre elles par la même thématique. Vu que tous

les résultats sont publiés avec des collaborateurs, les sous-chapitres sont introduits

par un petit résumé où je décris aussi ma contribution. Les sous-chapitres résultats

(1 à 4) sont des publications. Le sous-chapitre (5) a été soumis pour publication et

les deux derniers sous-chapitres (6 et 7) sont présentés, par soucis de concision,

très succinctement avec quelques figures. Ces deux derniers chapitres représentent

beaucoup de travail et sont en cours d’étude. La thèse se termine par une petite

conclusion et les références. Bonne lecture !

Page 11: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

2

Résumé de la thèse Introduction sur BARD1

Les cancers du sein et de l’ovaire sont prévalents dans les pays développés. En

effet, le risque cumulé dans la vie d’une femme de développer un cancer du sein est

de 11% aux Etats Unis. La plupart des cas sont sporadiques, et, seulement environ

10% sont d’origine héréditaire. Deux gènes principaux de susceptibilité aux cancers

gynécologiques ont été identifiés : BRCA1 et BRCA2 (Breast Cancer susceptibility

genes 1 and 2). Ces deux gènes sont responsables d’environ la moitié des cas de

tumeurs héréditaires. Des mutations dans d’autres gènes tels que PTEN, TP53,

ATM, FANCD2 peuvent aussi prédisposer au cancer du sein, mais avec un risque

très faible. BRCA1 et BRCA2 sont décrits comme des suppresseurs de tumeurs ; en

effet, des mutations dans ces gènes prédisposent à des tumeurs du sein et de

l’ovaire et les cellules de ces tumeurs perdent leurs allèles de type sauvage,

produisant une perte de fonction. La nécessité d’étudier les mécanismes de

suppression de tumeurs contrôlés par BRCA1 et de découvrir d’autres facteurs

génétiques prédisposant aux cancers gynécologiques, a mené à la découverte de

BARD1 (BRCA1 Associated RING Domain 1). Un criblage deux-hybrides dans la

levure, en utilisant le bout N-terminal de BRCA1 comme appât, a permis

l’identification d’une protéine interagissant avec BRCA1. Cette protéine, BARD1, est

aussi un suppresseur de tumeur, car des mutations dans son gène semblent

prédisposer au cancer du sein et parce que son absence n’est pas viable à l’état

embryonnaire.

La structure de BARD1

BARD1 est une protéine de 777 acides aminés chez l’homme et de 765 chez la

souris. BARD1 est un homologue de BRCA1. Elle contient deux domaines

homologues hautement conservés pendant l’évolution. Ces domaines sont le

domaine RING finger en N-terminal et le domaine BRCT en C-terminal. De plus,

BARD1 contient trois domaines ankyrines que l’on ne se retrouve pas chez BRCA1.

BARD1 et BRCA1 interagissent à travers leurs domaines RING. Deux chaînes en

hélice-α de part et d’autre de la séquence primaire du RING assurent les liaisons

hydrophobes permettant l’assemblage des deux domaines RING entre eux. Le RING

Page 12: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

3

a une structure conservée qui contient deux atomes de zinc, nécessaires à l’activité

enzymatique de l’hétérodimère BRCA1/BARD1. BARD1 a plusieurs orthologues chez

d’autres espèces comme chez le Xenopus laevis et le nématode Caenorhabditis

elegans, avec une fonction conservée.

Les modèles génétiques

Pour comprendre la fonction de BARD1, une analyse génétique a été effectuée dans

des souris. Comme pour BRCA1 ou BRCA2, les souris knock-outs pour BARD1 ne

sont pas viables et meurent toutes au stade embryonnaire. Les embryons meurent

prématurément parce que les cellules ne prolifèrent plus normalement et pas à cause

de la mort cellulaire par apoptose. En revanche, les souris possédant un knock-out

inductible développent des tumeurs du sein au bout d’une année. De plus, les

cellules avec une délétion du gène de BARD1 génèrent rapidement une instabilité

chromosomique et une morphologie néoplasique. Ces résultats montrent

l’importance de l’expression de BARD1 dans le maintien du génome et dans la

suppression de cancers.

L’activité ubiquitine ligase de BARD1

La fonction biochimique principale de BARD1 réside dans sa capacité à ubiquitiner

des protéines. Cette fonction E3 ubiquitine ligase est augmentée en présence de

BRCA1. Bien que des homodimères de BARD1 et de BRCA1 aient été observés in

vitro, seule la forme hétérodimèrique est fonctionnelle in vivo. La fonction ubiquitine

ligase de l’hétérodimère est nécessaire dans la suppression tumorale. En effet, les

mutations connues, telles que C61G et C64G, inhibant l’activité ubiquitine ligase du

domaine RING prédisposent à des tumeurs gynécologiques. Ces mutations

déstabilisent la structure tridimensionnelle contenant le zinc, mais ne détruisent pas

les liaisons hydrophobes de l’hétérodimère. Le domaine RING est caractéristique des

ubiquitine ligases. Ces ubiquitine ligases sont très utiles, voire indispensables, dans

le métabolisme de protéines contrôlant la progression du cycle cellulaire, telles que

des cyclines D1, E1, p27, etc., et de ce fait peuvent jouer un rôle important dans le

cancer et sa suppression. Les cibles de l’ubiquitination de BARD1/BRCA1 sont peu

connues, probablement importantes dans la régulation du cycle cellulaire, mais on

sait que BARD1/BRCA1 peut s’auto-ubiquitiner. De plus, BARD1/BRCA1 peut

ajouter une protéine d’ubiquitine sur les histones H2A et H2AX, et sur FANCD2. Une

Page 13: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

4

autre cible de l’ubiquitination de BRCA1/BARD1 est la γ-tubuline ce qui est important

dans la régulation de la duplication des centrosomes. En outre, la

nucleophosmine/B23 est aussi une cible candidate pour une ubiquitination par

BRCA1/BARD1, confirmant un rôle pour BARD1 dans la régulation des centrosomes.

Finalement, l’hétérodimère BRCA1/BARD1 peut se lier la RNA polymérase II

(RNAPII) et peut ubiquitiner la sous-unité Rbp1 de RNAPII.

Les fonctions de BARD1 dépendantes de BRCA1

BARD1 a plusieurs fonctions. BARD1 se lie à BRCA1, et semble partager beaucoup

de fonctions qui sont attribuées à BRCA1. Cependant, BARD1 possède des

fonctions indépendantes de BRCA1. Sa fonction la plus importante est partagée

avec BRCA1 dans la réparation de l’ADN. BARD1 interagit avec BRCA1 dans la

réparation de l’ADN cassé en doubles brins (DSB, double strand breaks) et est

activée par ATM, ATR et en se liant au DNA-PKs et aux protéines Ku70 et 80. La

réparation peut se faire par HR (homology directed repair), par NHEJ (Non-

homologous end joining), ou par MMR (mismatch repair) en interagissant avec

hMSH2, hMSH3, hMSH6. L’expression de BARD1 peut être régulé pendant le cycle

cellulaire, mais pour la réparation de l’ADN, BARD1 est localisé dans des complexes

nucléaires pendant la phase S avec BRCA1 et Rad51 au niveau des sites de

fourches de réplication et peut se re-agréger dans des complexes de réparation

d’ADN avec PCNA et Rad51 après traitement hydroxyurée. La localisation nucléaire

n’est pas exclusive et le déplacement de BARD1 dans le cytoplasme sera discuté

dans la partie résultats.

BARD1 et BRCA1 jouent un rôle dans la régulation de la transcription car elles

peuvent se lier au complexe RNA polymerase II holoenzyme (holo-pol). En outre,

holo-pol est dégradée par ubiquitination après dommage de l’ADN. L’arrêt de la

transcription à des sites de dommages d’ADN est associé à la dégradation de holo-

pol suggérant un rôle pour BRCA1/BARD1 ubiquitine ligase dans la régulation de la

transcription. De plus, il semblerait que le domaine BRCT de BARD1 comme celui de

BRCA1 peut activer la transcription in vitro. BRCA1 peut se lier directement à l’ADN,

sans doute à des sites de fourches de réplication. BARD1, en outre, peut se lier avec

le facteur CstF-50 dans le but d’inhiber la polyadenylation et la maturation des ARN

messagers in vitro, indiquant que BARD1 a effectivement un rôle dans la régulation

Page 14: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

5

de l’expression de gènes. La sous-unité p50 de NF-kB peut se lier indirectement à

BARD1 par l’intermédiaire de Bcl-3. L’interaction Bcl-3-BARD1 pourrait avoir lieu au

niveau de leurs domaines ankyrines. Cette interaction est une indication

supplémentaire quant au rôle de BARD1 dans la régulation de la transcription de

certains gènes.

Les fonctions de BARD1 indépendantes de BRCA1

En plus des fonctions de BARD1 dépendantes de BRCA1, BARD1 a aussi une

fonction indépendante de BRCA1, car sa surexpression dans des cellules en cultures

est suffisante pour induire l’apoptose. Cette fonction est importante mais sera donc

discutée plus en détails dans la thèse.

Objectifs de la thèse

Le but de cette thèse est d’étudier les mécanismes de suppression tumorale

contrôlée par BARD1 et de comprendre comment BARD1 maintient une stabilité

génomique. La première partie de la thèse consiste à répondre aux questions

suivantes : quelle est la partie de BARD1 qui est nécessaire à l’induction de

l’apoptose ; quels sont les déterminants de cette voie apoptotique ; quels sont les

protéines qui interagissent avec BARD1 dans la suppression tumorale ; est-ce que

p53 interagit dans cette voie signalétique de l’apoptose? En outre, on a voulu

montrer la dynamique de BARD1 et déterminer son expression et sa localisation

pendant l’apoptose ainsi que pendant la mitose ; d’identifier les protéines qui

interagissent avec BARD1 pendant la mitose et qui régulent son expression en

caractérisant son promoteur. Dans un deuxième temps, nous avons étudié le

phénomène de l’instabilité chromosomique dans les cellules cancéreuses et identifié

certains déterminants de la stabilité génomique pendant la mitose, l’effet de BARD1

sur la régulation des télomères et sur la ségrégation des chromosomes.

Page 15: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

6

Résultats et discussion

1. BARD1 : un médiateur de l’induction de l’apoptose en réponse à un stress génotoxique, dépendant de p53 BARD1 et BRCA1 ont des fonctions communes dans la réparation de l’ADN, dans la

stabilité génomique, ainsi que dans des fonctions antitumorales. En effet, ces deux

protéines homologues sont liées par leurs domaines RING respectifs. Dans une

étude précédente, nous avons montré que la répression de BARD1 provoque des

répercussions au niveau de la stabilité chromosomique ainsi que sur la morphologie

et physiologie cellulaires, ce qui suggère une évolution vers un phénotype

néoplasique. Dans cette première publication, nous montrons que BARD1 peut avoir

des fonctions indépendantes de BRCA1 ainsi que l’importance de BARD1 face à un

stress endommageant l’ADN. En cas de stress cellulaire et nucléaire, l’expression de

BARD1 est augmentée à l’instar de celle de p53. Ces augmentations concomitantes

de BARD1 et de p53 ont été observées in vitro en culture cellulaire, et in vivo dans

des cas d’hypoxies induites par la formation d’une ischémie vasculaire cérébrale

chez la souris. Nous avons démontré que cette augmentation se fait au niveau de la

protéine et de son ARN messager. Il est intéressant de noter que les cellules qui sont

réprimées pour BARD1 sont moins aptes à entrer en apoptose quand elles sont

soumises à un stress. En effet, la fragmentation de l’ADN, normalement activée par

des endonucléases spécifiques, de même que la présence de la caspase-3 activée

sont réduites dans les cas où BARD1 est réprimée ou supprimée. La surexpression

de BARD1 par transfection induit une mort cellulaire programmée. Les cellules

produisent les mêmes signes qu’en cas de stress génotoxique, à savoir la

fragmentation de l’ADN évaluée sur gel d’agarose, et mesurée quantitativement par

cytométrie de flux (méthode TUNEL). Un mutant de BARD1 comprenant une

mutation ponctuelle Q564H est incapable d’induire l’apoptose. De plus, l’induction de

l’apoptose peut se faire même en l’absence de BRCA1. Les cellules souches de

souris homozygotes BRCA1-/- peuvent être également induites en apoptose. Ceci

démontre que BARD1 a une action cellulaire indépendante de BRCA1. Par contre,

les effets pro-apoptotiques de BARD1 sont dépendants de p53 active et

fonctionnelle. De plus, nous avons montré que BARD1 peut interagir avec p53

physiquement par des expériences de co-immunoprécipitations.

Page 16: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

7

2. La translocation de BARD1 du noyau vers le cytoplasme est liée à son activité apoptotique BARD1 est un suppresseur de tumeur ainsi qu’une protéine pro-apoptotique.

Puisqu’une augmentation de l’expression de BARD1 peut suffire à induire une

réponse apoptotique dépendante de p53, nous avons cherché à identifier le domaine

protéique de BARD1 responsable de la signalisation menant à l’apoptose. Pour cela,

une série de mutants de délétion de BARD1 a été transfectée dans des cellules en

culture et nous avons mesuré la capacité des protéines recombinantes à induire

l’apoptose. Une région de la séquence de BARD1 comprenant les acides aminés de

510 à 604 est nécessaire à l’induction de l’apoptose. Cette partie de la séquence de

BARD1 se trouve entre deux domaines fonctionnels : les répétitions d’ankyrines et le

domaine BRCT en C-terminal. Ce domaine ne correspond pas à une structure

conservée de la protéine. Ce domaine a la particularité de contenir les deux

mutations répertoriées dans les tumeurs du sein C557S et Q564H. La partie pro-

apoptotique de BARD1 corrèle donc avec la fonction de suppresseur de tumeur. De

plus, cette étude démontre que la localisation nucléaire de BARD1 n’est pas

exclusive, malgré la présence de nombreux NLS (signaux de localisation nucléaire).

En effet, des expériences d’immunofluorescence ou avec la protéine de fusion

BARD1-EGFP, et des protéines de délétion fusionnées avec EGFP ont montré que

BARD1 peut également être exporté dans le cytoplasme. La détection cytoplasmique

de BARD1 a pu se faire in vivo et in vitro : sur des sections de tissus de rate de

souris par immunohistochimie d’une part, et sur des cellules en culture par

immunofluorescence et par microscopie électronique d’autre part. Les extraits de

cytoplasme et de noyau montrent par western blot une présence de BARD1 dans le

cytoplasme qui augmente en fonction du temps après induction de l’apoptose. Cette

augmentation sensible de BARD1 dans le cytoplasme est accompagnée de la

présence d’une protéine tronquée de p67 qui est peut-être le résultat d’un clivage

protéolytique. La double localisation nucléaire et cytoplasmique de BARD1 dénote

une double fonction : nucléaire dans la réparation de l’ADN et cytoplasmique dans la

mort cellulaire programmée.

Page 17: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

8

3. L’expression de BARD1 pendant la spermatogenèse est associée à l’apoptose et peut être régulée hormonalement L’expression protéique de BARD1 est élevée dans les tissus germinaux males et

pendant l’embryogenèse. Dans cette étude, nous démontrons que BARD1 est en fait

exprimé à tous les stades de la spermatogenèse contrairement à BRCA1 qui est

seulement exprimé dans les spermatides ronds et les spermatocytes primaires

méiotiques. Alors que les spermatogonies expriment le messager complet de

BARD1, les spermatocytes précurseurs tardifs expriment une nouvelle forme de

BARD1, plus courte, qui est prédominante et qui est désignée BARD1β. Ce

messager codant pour BARD1 tronquée, dépourvue de son RING domaine, a perdu

sa capacité à se lier et d’interagir avec BRCA1. Par contre, BARD1β conserve son

aptitude à induire l’apoptose. Ce fait est important dans la mesure où les

spermatogonies qui prolifèrent rapidement ont besoin d’un système efficace pour

éliminer les cellules surnuméraires par apoptose. La mort cellulaire par le truchement

de BARD1, permet cette régulation. De plus, il a été démontré que la forme tronquée

de BARD1β peut induire l’apoptose in vitro par transfection transitoire de cellules en

culture. L’apoptose a été suivie in vivo par immunohistochimie sur des sections de

testicules de souris. L’expression de BARD1 est corrélée avec l’expression de BAX

et de la caspase-3 active ainsi que la fragmentation de l’ADN. Finalement, nous

avons montré dans cette étude que, dans des cas de pathologies testiculaires telles

que l’azoospermie obstructive, la cryptorchidie et le syndrome de Sertoli, l’expression

de BARD1 est augmentée.

4. BARD1 induit l’apoptose en catalysant la phosphorylation de p53 par une kinase activée en réponse aux dommages de l’ADN Il a déjà été démontré que BARD1 joue un rôle dans la réponse apoptotique

indépendamment de BRCA1, mais cela nécessite la présence de p53. Sans p53, la

surexpression de BARD1 ne peut pas induire l’apoptose dans des cultures de

cellules transfectées. Cette thèse est renforcée par le fait que BARD1 peut co-

immunoprécipiter p53 et le stabiliser. L’activation de p53 pendant l’apoptose se fait

par l’ajout post-traductionnel de groupements phosphates sur plusieurs résidus,

notamment sur les serines 15 et 37. Les kinases identifiées à ce jour dans la réponse

aux dommages de l’ADN sont multiples et se fait par ATM, ATR et DNA-PK. Dans

Page 18: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

9

cet article, nous montrons que BARD1 peut catalyser la phosphorylation de p53 sur

la serine 15. Cette phosphorylation est effectuée par une kinase, qui a donc besoin

de la présence de BARD1 fonctionnel. Elle est sans doute effectuée à travers le

rapprochement et l’attachement de BARD1 à p53. Pour démontrer que l’interaction

est nécessaire nous avons co-immunoprécipité p53 avec des mutants de délétions

fusionnés à la EGFP permettant la reconnaissance de la protéine recombinante. Les

divers mutants nous indiquent que la zone d’attachement de BARD1 à p53 est

relativement importante, puisqu’elle couvre pratiquement toute la séquence primaire

de BARD1. Néanmoins, cette expérience nous montre que le mutant qui ne contient

pas la séquence codant pour le RING domaine, ainsi que la EGFP seule, sont

incapables de co-immunoprécipiter p53. Ce résultat nous indique également que

BARD1 se lie à p53 sans le soutient de BCRA1. Dans cette étude, nous montrons

que BARD1 peut aussi co-immunoprécipiter Ku-70, une kinase impliquée dans la

reconnaissance de cassure d’ADN double brin. Cette kinase est activée après

exposition à un stress nucléaire et cellulaire et elle peut être la kinase responsable

de la phosphorylation de la serine 15 de p53 induite par BARD1, puisque BARD1

peut se lier à elle.

5. BARD1 interagit avec BRCA2 afin de réguler la progression du cycle cellulaire BARD1, BRCA1 et BRCA2 sont des suppresseurs de tumeurs qui permettent le

maintien de l’intégrité des chromosomes pendant la mitose. La ségrégation des

chromosomes doit se faire sans perte ou gain sinon la résultante est une instabilité

génomique. Dans les souris knock-out, la suppression des gènes susmentionnés

provoque une mort embryonnaire précoce due à des problèmes de prolifération

cellulaire et non à l’apoptose. Les cellules dépourvues de BARD1, BRCA1 ou

BRCA2 par knock-down siRNA ont une grande propension à devenir instables

génétiquement, la marque universelle des cellules tumorales. Cette étude vise donc

à identifier les déterminants et les liens entre la suppression tumorale réglée par

BARD1 et la maintenance de l’instabilité chromosomique dans les cellules en

division. En utilisant un siRNA de BARD1 pour réduire l’expression de son messager

et par conséquent de son produit protéique, nous observons une augmentation de

temps du cycle cellulaire, sa complétion est donc ralentie. De surcroît, les cellules

déficientes en BARD1 présentent des anomalies dans la cytokinèse. En examinant

Page 19: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

10

l’état du génome, après analyse du caryotype des cellules transduites avec le siRNA

de BARD1, nous avons constaté l’apparition de cellules génétiquement instables

contenant des chromosomes surnuméraires ou présentant des pertes

chromosomiques. De plus, il a été démontré par nous et par d’autres groupes que

BARD1 varie pendant le cycle cellulaire et atteint un pic d’expression pendant la

mitose, soit en G2/M. Cette augmentation de BARD1 s’accompagne d’une co-

localisation partielle de BARD1 avec les microtubules du fuseau mitotique. Cette co-

localisation a été observée par microscopie à fluorescence. L’expérience étant basée

sur la spécificité de l’anticorps, les contrôles de spécificité de l’anticorps dans ces cas

sont importants. Nous avons montré que l’anticorps anti-BARD1 choisi réagit

uniquement avec la protéine de fusion BARD1-EGFP et non avec les protéines

tronquées en N-terminal, ce qui défini la région de l’épitope. Pour comprendre les

mécanismes de BARD1 dans la cytokinèse, nous avons voulu savoir si BARD1

pouvait interagir avec les kinases qui sont impliquées dans la cytokinèse.

Effectivement, BARD1 co-immunoprécipite Aurora B kinase et non Aurora A. De

plus, et à notre surprise, BRCA2 peut co-immunoprécipiter avec BARD1 seulement

dans la phase G2/M au moment de la phosphorylation de BRCA2. Ces résultats

nous donnent une indication supplémentaire sur le rôle de BARD1 dans la régulation

du cycle cellulaire. Les voies signalétiques de cette régulation ne sont néanmoins

pas encore connues, bien que la fonction ubiquitine ligase de l’hétérodimère peut

être mise en cause.

6. Le rôle de BARD1 dans le maintien des télomères Nous savons que la majorité des cellules cancéreuses ont une instabilité

chromosomique ainsi qu’une télomèrase active. Nous avions précédemment vu que

la non expression de BARD1 dans des cellules en culture ou bien dans des souris

knock-out produit une augmentation de l’instabilité génomique. De plus, il a été

démontré que BARD1 et BRCA1 sont des facteurs qui s’agrègent dans les corps

nucléaires, PML, dans les cellules négatives pour la télomèrase et que le complexe

BRCA1-Mre11-Rad50-Nbs1, qui intervient dans la réparation d’ADN, est aussi

présent dans le complexe télomèrique. En sachant que BARD1, tout comme les

télomères, s’assemblent en agrégats nucléaires, et que les protéines télomèriques

sont nécessaires dans le maintien de l’intégrité des chromosomes, nous avons voulu

déterminer s’il y avait un lien de causalité entre un dysfonctionnement de BARD1,

Page 20: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

11

des télomères et une instabilité chromosomique. En d’autres termes, est-ce que

BARD1 peut moduler le fonctionnement des protéines télomèriques et le maintien

des télomères? Nous avons voulu savoir si BARD1 avait une influence fonctionnelle

sur la longueur des télomères, sur l’activité de la télomèrase et si BARD1 pouvait se

localiser avec les structures télomèriques. Les résultats d’immunohistochimie avec

des sondes télomèriques et l’anticorps dirigé contre BARD1 sur des cellules en

cultures et sur des sections de tissus nous montrent qu’une fraction de BARD1 a

tendance à se localiser sur ou près des régions télomèriques. De plus, il apparaît

dans certains cas que BARD1 peut co-localiser avec TRF-2 une protéine couvrant la

longueur des télomères. Afin de voir s’il y avait une interaction directe entre BARD1

et les télomères, nous avons tenté d’immunoprécipiter l’ADN télomèrique avec un

anticorps anti-BARD1. Le résultat obtenu par chromatine IP (ChIP) s’est révélé peut

concluant ne démontrant pas d’interaction forte entre BARD1 et la chromatine

télomèrique. La longueur des télomères des cellules épithéliales de glande

mammaire de souris, qui ont été réprimées pour BARD1, a été mesurée par des

méthodes de flow-FISH et de TRF. Il semblerait que ces cellules, comparées aux

cellules normales, aient subit dans certains cas un raccourcissement des télomères.

En fait, la longueur des télomères devient surtout beaucoup plus hétérogène dans

les cellules BARD1 réprimées, ce qui est peut être l’expression d’un dérèglement du

maintien des télomères. Puisque l’inactivation de BARD1 peut induire un phénotype

néoplasique, et que la majorité des cellules tumorales ont une réactivation de la

télomèrase, nous avons voulu savoir si la diminution de l’expression de BARD1 avait

une influence négative sur l’activité de la télomèrase. Nous avons donc mesuré

l’activité de la télomèrase par un test enzymatique et par PCR (TRAP) dans les

cellules où l’expression de BARD1 était réduite. Ces expériences n’ont pas apporté

de résultats concluants. En effet, les cellules de glandes mammaires de souris

utilisées sont négatives pour la télomèrase à l’état sauvage. Les prochaines

expériences tenteront de voir s’il y a une diminution de la longueur des télomères

dans des lignées de cellules primaires et cancéreuses où on a réduit l’expression de

BARD1 par transduction lentivirale de siRNA. La mesure de la diminution de la

longueur des télomères devra s’effectuer sur plusieurs générations par la méthode

dite de « Southern », par hybridation d’une sonde télomèrique marquée sur un gel

d’ADN télomèrique (méthode TRF, telomere restriction fragment assay). Les

résultats sont préliminaires et ne permettent pas de statuer sur une fonction de

Page 21: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

12

BARD1 sur les télomères. Il est possible qu’en dépit d’un effet de BARD1 sur les

télomères, que son action soit indirecte, un épiphénomène.

7. Caractérisation du promoteur de BARD1 Nous avons déjà constaté que BARD1, au niveau de son messager et de sa

protéine, est augmentée dans des cas de stress génotoxique et qu’il est régulé

hormonalement. De plus l’expression et la stabilité de BARD1 et de BRCA1 sont

interdépendantes. Afin d’étudier les mécanismes de régulation d’expression de

BARD1 dans les diverses conditions de stress cellulaire, de tumeurs, de réponse à

une exposition aux hormones, de son rôle dans l’inflammation, de mort cellulaire, ou

même dans le développement embryonnaire, il était important d’isoler et de

caractériser le promoteur de BARD1 et la région promotrice minimale requise pour

répondre à tous ces stimuli. Nous avons donc, dans un premier temps, caractérisé la

région initiatrice de la transcription, le site cap en 5’ en amont de l’exon 1. Par une

expérience 5’RACE et par PCR nous avons défini le site cap en amont de ce qui a

déjà été publié. Nous avons amplifié par PCR à partir d’ADN génomique humain une

région promotrice supposée que nous avons cloné dans un vecteur contenant le

gène reporteur de la luciférase. Après transfection dans différentes lignées de

cellules animales, nous avons constaté que la séquence de 856 paires de bases en

amont du 5’UTR est suffisante pour l’activation constitutive de BARD1, mais en

revanche ne semble pas contenir de sites de liaison à d’éléments régulateur. En

effet, après une exposition à un stress cellulaire ou un traitement hormonal, les

cellules transfectées avec le gène reporter ne montrent pas de variation notoire dans

l’expression de la luciferase. Néanmoins, des travaux sont en cours pour caractériser

une région qui pourrait réguler sensiblement l’expression du gène de BARD1.

Notamment, une région de 2402 paires de bases en amont du 5’UTR a été clonée

pour effectuer des tests par le système du gène reporter. Cette région contient des

sites putatifs d’attachement à la p53 et au NF-κB entre autres, qu’il convient

d’évaluer par des « shifts assay », « super shifts » et du DNA « footprinting » par

exemple, en présence ou non de facteurs stimulants tels que des cytokines ou du

TNF-α. Il serait intéressant de voir si cette région promotrice peut être activée par

des cytokines, puisqu’on sait que la surexpression de BRCA1 fait augmenter la

production cellulaire de NF-κB et que la sous-unité p50 peut se lier à BARD1 par

Page 22: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

13

l’intermédiaire de Bcl-3. Ce serait une voie possible de régulation de l’expression de

BARD1 par BRCA1, puisque l’on sait que BARD1 et BRCA1 se stabilisent

mutuellement.

Conclusion

Cette thèse divisée en plusieurs points se concentre sur une même thématique : le

rôle de BARD1 dans la suppression de tumeurs et dans le maintien de la stabilité

chromosomique et génétique d’une manière générale. Nous montrons que BARD1

agit souvent dans l’ombre de BRCA1, mais qu’elle peut avoir des fonctions

indépendantes de BRCA1 dans plusieurs processus cellulaires. BARD1 possède une

fonction proapoptotique. Bien que BARD1 soit décrite comme une molécule

nucléaire, elle est également une protéine dynamique pouvant se localiser dans le

cytoplasme. De plus, la fonction apoptotique est corrélée à sa localisation

cytoplasmique. En outre, BARD1 joue un rôle dans l’apoptose des spermatocytes

surnuméraires. BARD1 se lie à p53, la stabilise et est impliquée dans sa

phosphorylation. Elle régule la complétion du cycle cellulaire, par TACC1, Aurora B

et BRCA2 et semble avoir un rôle dans le maintien de l’intégrité des télomères.

BARD1 est donc une protéine multifonctionnelle qui doit encore nous révéler bien

des secrets au niveau de ses voies signalétiques et de ses fonctions.

Page 23: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

14

Preface

The purpose of cancer

Question. Cancer is one of the leading causes of death in the world. In fact, it is the

second leading cause of death after cardiovascular diseases. It has been predicted

that one person in four will be afflicted with cancer during their life time. It is without

any doubt a terrible and devastating disease. However, for a Doctor of Philosophy,

that should not be the only reason for studying cancer. Compellingly, tumour

development raises profound questions about life and death, and particularly, on the

proliferation, survival and death of a single cell. So what is cancer, how does it occur

and what purpose does it serve?

What. Cancer is a paradox: it is life and death at the same time. Cancer is life

because it is nothing else but the story of a cell escaping the sentence of death and

becoming immortal. The cancer cell will divide and make a clone. It will create a

population of cells that have acquired non-stop cell division. Cancer cells, like

unicellular organisms, behave selfishly. The never ending self-replication is the

default survival mechanism of all unicellular organisms, a mechanism without which

life on earth would not be possible. However, in higher organisms, homeostasis and

renewal of complex tissues requires regulated cell proliferation. Regulated cell

proliferation means that cell division needs to be flawlessly counter-balanced by

apoptosis, i.e. programmed cell death. In essence, the number of dividing cells

should equal precisely the number of eliminated cells. Disrupting that equilibrium with

a shift towards cell survival and uncontrolled proliferation is cancer. Unfortunately,

neoplasia, or malignant cellular transformation, inexorably ensues in death, in the

decease of its host. Therefore, cancer is also the expression of death.

How. A large body of evidence has demonstrated that evolution is a consequence of

genetic mutability in germ cells followed by a selection process under the pressure of

the environment. Consequently, genetic variability emerging either in point mutations

or genome alterations during meiosis is the necessary driving force of evolution and

speciation. Likewise, tumour cells arise through a series of pair-wise mutations,

Page 24: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

15

clonal selection and waves of expansion, which may be considered as a

consequence of Darwinian selection. Therefore, cancer can well be regarded as an

evolutionary consequence of genetic instability in somatic cells. Although

environmental stress increases mutation rate, genetic instability, and cancer

incidence, tumour cells may acquire mutations at a normal rate. However, cancer

remains a naturally occurring genetic disease. Consequently, all cancers go through

a state of genomic instability, resulting in chromosomal alterations and in aneuploidy.

Why. Our own desire of eradicating cancer is, in a way, the expression of our

yearning for immortality, or perhaps more modestly, a healthy long life. Because

evolution requires the removal of old and deleterious genes and the renewal of the

gene pool, the fitness and survival of an entire species relies on the elimination of a

living organism at a pre-reproductive stage, but also at a post reproductive stage.

Furthermore, in order to avoid depletion of limited resources in a finite environment

by overcrowding and to provide an efficient turnover of generations, death of the

individual at a post-reproductive age is essential. Thus, if we humans could be

immortal, we would be considered as cancers, because we would be a liability to our

natural environment. In other words, programmed cell death is a necessity for the

survival of the whole organism, such as the death of an individual is a necessity for

the perpetuation of the species and its evolution.

Solution. Because mutations occur naturally over time, cancer incidence increases

with age. Coincidently, the process of aging is also a by-product of genetic instability,

just like cancer. Therefore, if we humans lived long enough, we would all eventually

die of cancer, unless we genetically acquired “anti-cancer” mutations. Since death of

a human being is a prerequisite for the survival of the human race and that there is

no selective pressure preventing death, cancer would ensure, over time, death of an

individual. Finally, it could simply mean that cancer is a necessity maintained through

natural selection, in order to provide speciation and survival. In this case, cancer

would be the ultimate barrier to an extended lifespan in a disease-free world.

Question. Is cellular immortality a natural suicidal mechanism of an individual?

Page 25: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

16

Review on BARD1 and its diverse functions

I. Breast cancer and the BRCA genes

BRCA1 and BRCA2 mutations in cancer

Breast and ovarian cancers are highly prevalent in western societies. For instance, in

the United States, the cumulative risk of developing breast cancer is 11% of women

by the age of 80 (Casey et al., 1997)1. Most cases of breast and ovarian cancers are

sporadic since only about 10% are hereditary (Wooster and Weber, 2003)2. Less

than half of the familial breast and ovarian cancers, carry germ-line mutations in the

breast cancer susceptibility gene 1 or 2 (BRCA1 or BRCA2) (Hall et al., 1990; Futreal

et al., 1994; Miki et al., 1994, Wooster et al.,1994, 1995)3-7 (Figure 1).

Other known inherited mutated genes in familial breast and ovarian cancers include

PTEN (Shugart, 1999)8, and CHEK2 (Meijers-Heijboer, 2002; Schutte et al., 2003;

Caligo, 2004)9-11. Mutations in other genes such as TP53, associated with Li-

Fraumeni syndrome, ATM in Ataxia telangectasia, FANCD2,which is inactivated in

Breast and Ovarian Cancer Genes

other genesBRCA1 20%BRCA2 20%CHEK2 5%TP53 <1%PTEN <1%STK11/LKB1 <1%ATM <1%FANCD2 <1%

Figure 1. Representation of susceptibility genes predisposing to breast and ovarian tumours. Only half of the

hereditary genes have been discovered, the other half is probably polygenic and could have environmental

and multifactorial origins. These genes represent only 10% of all breast and ovarian cancers. The rest are

sporadic cases. (Data compiled from Wooster and Weber, 2003).

Page 26: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

17

Fanconi anaemia, and STK11/LKB1 mutated in cases of Peutz-Jeghers syndrome,

are also associated with breast cancer, but predispose at low frequencies (Smith et

al., 1993; Ford et al., 1995; Chen et al., 2000)12-14. The other familial breast and

ovarian cancers do not have an apparent autosomal dominant inheritance pattern

and are probably polygenic with environmental influences (Mincey, 2003)15.

BRCA mutations are highly penetrant, since women harbouring mutations in either

BCRA1 or BRCA2 have a 80-90% life-time risk of developing tumours in the breast or

in the ovaries, and account for about 40-50% of early onset of breast and ovarian

cancers (Easton et al., 1993,1994; Ford et al., 1994, Rahman and Stratton, 1998)16-19.

Moreover, in tumours from patients with inherited mutations in either BRCA1 or

BRCA2, the mutation of one allele is typically associated with the loss of their wild-

type allele, incurring a loss of function. This loss of function, also termed loss-of-

heterozygosity (LOH), means that BRCA1 and BRCA2 act as potent tumour

suppressors (Futreal et al., 1994)4 and correlates with the occurrence of breast and

ovarian cancer (Ford et al., 1998)20. Moreover, BRCA1 acts as a tumour suppressor

because its overexpression in tumour cells injected in nude mice induces growth

retardation of the tumour cells (Holt et al., 1996)21. Some sporadic breast and ovarian

tumours show defects in BRCA1 and BRCA2 genes or in their activity in tumour

tissues. This underpins the importance of epigenetic silencing pathways in breast and

ovarian tumours, adding another factor of complexity to BRCA-related tumours.

BRCA1 is a highly ubiquitous protein as it is expressed in almost all tissues. However,

tumours arise only in the breast and ovaries that have lost the remaining healthy wild-

type allele of individuals carrying germ-line mutations in the BRCA1 gene. LOH

occurring in other tissues has not been reported. Possibly because nullizygous Brca1

mutation in mice is non-viable for a developing embryo (Hakem et al., 1996; Gowen et

al., 1996)22,23. In these heterozygous individuals, other tissues are not affected and do

not develop tumours, and they do not appear to be able to lose their wild-type allele.

This conundrum, discussed in numerous reviews on BRCA1 (Monteiro, 2003; Starita

and Parvin, 2003; Venkitaraman, 2002, 2004)24-27, raises fundamental questions on

the function of BRCA1 and its differential activity in breast and ovaries. One may ask

why and how breast and ovarian tissues bear this permissive BRCA1 dysfunction

compared to other tissues.

Page 27: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

18

Genetic models for BRCA1 and BRCA2

To assess the function of BRCA1 and BRCA2 genes in breast cancer, knockout

mouse models were generated. The knockout mice, Brca1-/- or Brca2-/-, are not

viable and die early in embryogenesis. Several groups have shown that mouse

embryos with homozygous deletion in the Brca1 gene, die early in embryogenesis,

between days E5.5 and E13.5 depending on the type of deletion effectuated (Gowen

et al., 1996; Hakem et al., 1996; Liu et al., 1996; Ludwig et al., 1997)22,23. Such mice

typically showed several defects in gastrulation and in cellular proliferation. The

Brca2 knockout mice, carrying a homozygous truncation of exon 11 in Brca2,

displayed embryonic death at days E8.5-9.5 (Hakem et al., 1998)28. Mice which were

heterozygous for either Brca1 or Brca2 developed normally and were not more

susceptible to breast cancer than the normal wild-type littermates. However, mice

carrying conditional deletion knockout mutants for Brca1, driven under a mammary

tissue specific promoter, displayed mammary gland tumours by 10-13 months of age

(Xu et al., 1999a)29.

Overview of BRCA1 and BRCA2 functions

The structure and functions of BRCA1 and BRCA2 as well as the mechanisms-of-

action of BRCA-driven tumours have been widely discribed in numerous reviews

(Zheng et al., 2000; Scully and Livingston, 2000; Wang et al, 2000; Kerr and

Ashworth, 2001; Mueller and Roskelley, 2002; Moynahan, 2002; Narod and Foulkes,

2004; Venkitaraman, 2004)27,30-36. The BRCA1 and BRCA2 proteins interact with a

large number of proteins. They have a wide spectrum of activities since they play

roles in a variety of different pathways, coordinate and integrate various pathways,

which include DNA repair pathways and DNA replication cycle on one hand, and

centrosome duplication and mitotic checkpoints on the other.

BRCA1 is a ubiquitously expressed protein of 1863 amino acids consisting of two

highly conserved domains: a RING finger domain and two BRCT domains. Without

going into the detail of the multitude of BRCA1 functions, BRCA1 is involved in

diverse pathways such as DNA damage repair of double stranded breaks (DSB),

transcription-coupled repair, transcriptional regulation, chromatin remodelling,

ubiquitination, centrosome duplication and cell-cycle checkpoints (Deng and Brodie,

2000; Khanna, 2001; Starita and Parvin, 2003; Venkitaraman, 2004; Narod and

Page 28: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

19

Foulkes, 2004; Deng, 2002; Hsu and white, 1998; Hsu, 2001; Lotti et al.,2002;

Schlegel et al., 2003)25,27,36-41. One hypothesis could be that BRCA1 serves as a

scaffolding protein for the assembly of a multi-protein complex that plays a role in

integrating the aforementioned mechanisms.

The BRCA2 gene codes for a large protein of 3,418 amino acids, highly conserved

throughout evolution, and widely expressed in adult and embryonic tissues (Rajan et

al., 1997)42. It is structurally and phylogenetically unrelated to BRCA1 and bears no

homology with amino acid sequences of BRCA1 whatsoever. The function of BRCA2

is less diverse than that of BRCA1. Like BRCA1, BCRA2 is specifically involved in

homologous-directed repair (HDR or HR) and in non-homologous end joining

(NHEJ). BRCA2 attaches to Rad51 through its BRC repeats and binds directly to

DNA domains at sites of double stranded breaks (DSB).

To summarize, all the functions of BRCA1 and BRCA2 converge towards one goal

which is to maintain genome integrity. Since BRCA1 and BRCA2 are involved in DSB

repair, the common phenotypical feature of BRCA1 and BRCA2-deficient cells is that

they have a hypersensitivity to DNA damage-inducing agents such as ionising

radiation and inter-strand cross-linking drugs. These BRCA deficient cells display

increased genomic instability, as manifested by visible chromosomal aberrations and

aneuploid karyotypes (reviewed by Jasin, 2002; Moynahan, 2002)35,43. Chromosomal

aberrations are common to all, if not most, cancers and could be the pathogenic

basis of breast and ovarian cancers caused by mutations in BRCA1 or BRCA2.

However, investigating the function of these tumour suppressors, their associated

proteins and their molecular pathways is paramount in the fight over gynaecological

cancers.

II. BARD1 protein structure

BARD1 discovery and phylogenetic analysis

In the search for novel partners of BRCA1, BARD1 (BRCA1-Associated RING finger

Domain protein 1) was discovered in a yeast-two-hybrid screen using the N-terminus

fragment of BRCA1 containing the RING domain as bait (Wu et al., 1996)44.

Page 29: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

20

The human BARD1 gene is localized on chromosome 2 (2q34-q35) and the cDNA

encodes for a protein of 777 amino acids (Wu et al., 1996; Thai et al., 1998)44,45.

However, a doubt remains on the initiation site of translation, which could be situated

either at amino acid position 1 or 26. The BARD1 structure resembles BRCA1’s

structure, in that BARD1 shares sequence homology with BCRA1’s conserved

domains (Figure 2). Like BRCA1, the cloned human BARD1 contains a RING finger

domain at its N-terminus (residues 46-90) and two BRCT domains at the C-terminus

(residues 616-653 and 743-777). Moreover, BARD1, albeit less than half the size of

the BRCA1 protein (1863 residues), contains in addition three ankyrin domains

(ANK). Apart from the conserved domains, the rest of BARD1 bears no homology

with other proteins. BRCA1 is the only vertebrate protein that shares homology with

BARD1 in both the RING and BRCT domains. Homology in gene structure and

protein sequence of BARD1 and BRCA1 suggests that both proteins are

phylogenetically related. In contrast, BRCA1 and BRCA2 show a poor degree of

conservation through evolution. Indeed, mouse and human orthologues of BRCA1 or

BRCA2 share less than 60% of identical amino acid residues, but the functional

domains are conserved. The fact that the mouse Brca1 and Brca2 proteins share

mBARD1

Ankyrin repeatsRING Finger BRCT

9566 53 97 77 91 % of conservation with hBARD1

BRCA1

NLS 131 NLS 408 NLS 693

765 aa

1823 aa

BRCA2

777 aahBARD1

3418 aa

BRC repeats NLSTransactivation domain

(Not to scale)

mBARD1

Ankyrin repeatsRING Finger BRCT

9566 53 97 77 91 % of conservation with hBARD1

BRCA1

NLS 131 NLS 408 NLS 693

765 aa

1823 aa

BRCA2

777 aahBARD1

3418 aa

BRC repeats NLSTransactivation domain

(Not to scale)

Figure 2. Schematic of human and mouse BARD1, BRCA1 and BRCA2 proteins showing conserved

functional domains. BARD1 and BRCA1 are homologues since they share two conserved domains with high

homology, whereas BRCA2 is completely unrelated to either BARD1 or BRCA1.

Page 30: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

21

less than 60% homology with their human counterparts, contrasts with other known

tumour suppressors such as p53 or NF1, with 78% and 98% identity, respectively

(Connor et al.1997b)46. Comparison of amino acid homology between BARD1 and

BRCA1 shows that they had a common genetic ancestor and that they have co-

evolved at similar rates during evolution providing further clues that these proteins

may function in the same complex.

BARD1 orthologues

Alignment of human and mouse BARD1 orthologues shows low levels of amino acid

sequence homology, i.e. 70.4% over the total protein, whereas the RING, ANK and

BRCT motifs have 86.7% and 90.1% an 79.8% identity respectively (Irminger-Finger

et al., 1998; Ayi et al.,1998)47,48 (Figure 2). The fact that RING, BRCT and ANK

domains of BARD1 are highly conserved throughout evolution, and the rest of the

protein shows poor levels of homology, suggests that the three functional domains

mediate essential functions that are conserved through evolution. Furthermore, since

the isolation of BARD1 (Wu et al., 1996)44, orthologues from other species have been

cloned: Xenopus laevis (773 aa) (McCarthy et al., 2003), Rattus norvegicus (768 aa),

and Mus musculus (765 aa) (Figure 3).

Since then, other assumed orthologues from other species have been deduced from

their sequence homologies by reciprocal BLAST analysis. A predicted BARD1

Gallus gallus (chicken)

Gallus gallus (chicken)

Xenopus laevis (toad)

Rattus norvegicus (rat)

Mus musculus (mouse)

Takifugu rubripes (fugu fish)

Caenorhabditis elegans (nematode)

Pan troglodytes (chimpanzee)

Arabidopsis thaliana (thale cress)

Homo sapiens (human)

similar domain architectures

777 aa

765 aa

768 aa

772 aa

670 aa

734 aa

602 aa

1548 aa

750 aa

1083 aa

Gallus gallus (chicken)

Gallus gallus (chicken)

Xenopus laevis (toad)

Rattus norvegicus (rat)

Mus musculus (mouse)

Takifugu rubripes (fugu fish)

Caenorhabditis elegans (nematode)

Pan troglodytes (chimpanzee)

Arabidopsis thaliana (thale cress)

Homo sapiens (human)

similar domain architectures

777 aa

765 aa

768 aa

772 aa

670 aa

734 aa

602 aa

1548 aa

750 aa

1083 aa

777 aa

765 aa

768 aa

772 aa

670 aa

734 aa

602 aa

1548 aa

750 aa

1083 aa

Figure 3. Representation of known BARD1 orthologues in other species with conserved identical domains

showing similar domain architectures (source: http://www.ncbi.nlm.nih.gov/entrez/)

Page 31: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

22

hBRCA1

hBARD1

ANKYRIN REPEATS

RING DOMAIN BRCT

DOMAIN

BRCT DOMAINA

B

C

hBRCA1

hBARD1

ANKYRIN REPEATS

RING DOMAIN BRCT

DOMAIN

BRCT DOMAINA

B

C

orthologue encoding two of the three motifs exists in Arabidopsis thaliana (Joukov et

al., 2001; Irminger-Finger et al., 1998)47,49. The Caenorhabditis elegans BARD1

orthologue has been identified and conserves the same function as human BARD1

(Boulton et al., 2004)50. The chimpanzee orthologue has an overall homology of 99%,

but lacks the C-terminus and the two BRCT domains (Figure 3). In addition to the

three functional domains, human BARD1 contains 6 predicted nuclear localizing

signal (NLS) sequences, situated in the vicinity or embedded in each of the three

functional domains, inferring a predicted nuclear localization of BARD1. In

comparison to human BARD1, mouse BARD1 has only three NLS sequences with a

marginally lower prediction score for nuclear localization (Jefford et al., 2004)51.

Figure 4. Solution structure of the N-terminal and C-terminal domains of BARD1 and BRCA1. The RING

finger domain permits the interaction of BARD1 and BRCA1 creating a 4 alpha-helix bundle, held together

by hydrophobic bonds. The RING domain is flanked by the two alpha helices in the primary structure. The

RING motif contains the zinc binding domain. The two BRCT domains pack together forming a compact

structure (Modified from Baer; 2001, with reproductions from Brzovic et al., 2001; Williams et al., 2001).

Page 32: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

23

RING finger domain

The most studied domain in BARD1 and in BRCA1 is the RING finger domain. The

RING finger protein sequence motif was discovered 15 years ago and stands for

Really Interesting New Gene (Borden et al., 1998)52. The RING sequence motif is

defined as a unique array of conserved cysteine and histidine residues found at the

extreme N- or C-terminus of proteins that are typically involved in ubiquitination

pathways (Freemont, 2000)53. RING finger domains mediate also protein-protein

interactions and are involved in supra-molecular self-assemblies (Kentsis and Borden

2004)54. The conserved RING motif of BARD1 includes residues 50 to 86 and

corresponds to residues 24 to 64 in BRCA1 (Figure 4b). The RING finger binds two

atoms of zinc in a unique “cross-brace” conformation. The spacing between the

second and third pairs of cysteine and histidine residues implies that the distance is

conserved between the two zinc binding sites. BARD1 is the main binding partner of

BRCA1 and both proteins heterodimerize through their N-terminal RING finger

domains to form the active enzymatic complex (Figure 4a). The heterodimeric

interaction is strong, and although BRCA1 and BARD1 homodimers may be

produced in vitro, it is usually the heterodimers that are found in vivo (Meza et

al.,1999)55. Limited proteolysis analysis has defined that the region necessary for

heterodimerization includes residues 1-109 of BRCA1 and 26-119 of BARD1 (Meza

et al., 1999)55, (fig 5a). The cognate BARD1 and BRCA1 proteins heterodimerize by

forming an extensive 4-helix bundle each consisting of two anti-parallel alpha helices

and bind through their hydrophobic residues (Brzovic et al.,2001a; Morris et al.,

2002)56,57. The resolution of the crystal structure reveals that BARD1 has two alpha

helices, which flanks the RING motif in the primary sequence, and two short beta

sheets of the RING motif is separated by a loop which has the conserved pattern of

seven cysteine and one histidine residues which houses the two separate zinc-

binding sites (I and II) (Brzovic et al., 2001b)58 (fig. 4b). On the other hand, the

BRCA1 RING domain which has a similar structure contains three alpha helices and

two beta-sheets. RING finger domains are essential for the proper function of

proteins that contain them and for ubiquitination function. This is evidenced by the

multiple cancer predisposing mutations within the BRCA1 RING domain which

destroys BRCA1/BARD1 tumour suppressor and ubiquitin ligase activities (see

below). Several known cancer-predisposing mutations have been identified within

Page 33: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

24

this region. A missense mutation at either cysteine residue C39G, C64G, or C61G of

BRCA1 alters the Zn2+ binding site II and prevents the metal atom from binding to site

II. The site II cancer predisposing mutations C61G and C64G do not disrupt

heterodimer formation, but only the RING domain’s zinc binding capabilities (Morris

et al., 2002)57. Several residues have been identified within the alpha helices of

BARD1 that are required for heterodimer formation (Figure 5); these residues do not

participate in binding the zinc atom (Morris et al., 2002)57. Furthermore, site II cancer-

predisposing mutations in BARD1 RING domain have not yet been discovered,

possibly because the binding site for the E2/UbcH5 conjugating enzyme is localized

on the face of BRCA1 RING domain (Brzovic et al., 2003)59. It is rather mutations in

the hydrophobic interface that affect structural stability, even though some mutations

away from helical core are also capable of inhibiting heterodimer formation (Morris et

al., 2002)57.

Figure 5. Identification of residues of BARD1 required for the BRCA1/BARD1 heterodimer interaction. A)

Depiction of the positions of the amino acids which can inhibit heterodimer formation when mutated. The zinc

atoms are the grey spheres. B) Summary of mutations that inhibit heterodimer formation that were discovered

in a yeast two-hybrid screen. (Reproduced from Morris et al., 2002).

BARD1

BRCA1

A

B

BARD1

BRCA1

A

B

Page 34: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

25

Ankyrin repeats

Ankyrin domains are usually found in sets of repeats, usually of 3 or 4, but can have

up to 20 repeats. They mediate protein-protein interactions in very diverse families of

proteins, (for a review, refer to Mosavi et al., 2004)60. Their precise functions in

BARD1 remain elusive, but could be responsible for the numerous interactions

BARD1 may have with several classes of proteins implicated in tumour suppression.

BRCT domain

The BRCT domain (BRCA1-C-terminus) was first identified in BRCA1 (Koonin et al.,

1996). It is a conserved domain which is defined by distinct clusters of hydrophobic

residues folding into a unit of approximately 95 amino acids. The crystal structure of

XRCC1 BRCT domain has been elucidated and showed that pairs of these structures

may form homodimers as well as heterodimers (Zhang et al., 1998, Huyton et al.,

2000)61,62. The two BRCT domains of BCRA1 may bind together head-to-tail fashion

tethered by a poorly defined 23 amino acid linker (Williams et al., 2001)63 (Figure 4c),

forming a functional domain that can mediate transactivation functions and serve as a

multi-protein scaffold for several protein-protein interactions (Huyton et al., 2000)62.

BRCT repeats are found in many proteins involved in cell-cycle checkpoint pathways

responsive to DNA damage, DNA repair and recombination, such as 53BP1, RAD9,

RAD4, DNA ligase III and XRCC1 which bind to each other strongly (Ljunquist et al.,

1994; Caldecott et al., 1995)64,65. The BRCT domain of BRCA1 is essential for DNA

repair, transcriptional regulation, as well as several tumour suppressor functions (Hall

et al., 1990, Miki et al., 1994; Futreal et al., 1994; Friedman et al., 1994)3-5,66. BRCT

may bind to other BRCT repeats and other proteins with various unrelated structures.

Also, several studies have shown that the BRCT element of BRCA1 may bind or

interact with a number of proteins involved in transcriptional regulation, such as CtIP

co-repressor, histone deacetylases, p53, p300 and the DNA damage-associated

helicase BACH1 (Cantor et al., 2001; Deng and Brodie 2000)37,67. Furthermore, a

study showed that the tumour suppressor function of BRCA1 requires that the two

BRCT domains remain packed together (Huyton et al., 2000; Williams et al.,

2001)62,63. Two missense mutations within the BRCT domains of BRCA1 that were

found to be linked to breast and ovarian cancer (Miki et al., 1994, Futreal 1994)4,5

have been shown to impair correct BRCT homodimerization (Williams et al., 2001)63.

The tandem BRCT domains of BARD1 are predicted to behave in the same fashion,

Page 35: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

26

but its exact function still remains to be elucidated (Huyton et al. 2000)62. It would be

interesting to see whether the BRCT domains of BARD1 and BRCA1 interact.

III. Ubiquitin ligase function of BRCA1/BARD1

RING finger domain proteins are ubiquitin ligases

Many proteins containing a RING domain exhibit ubiquitin (Ub) ligase activity (E3)

(Freemont, 2000). The BRCA1/BARD1 heterodimer is an E3 ubiquitin ligase. There

are 890 genes in Homo sapiens that contain a RING domain and 213 are known to

be involved in ubiquitin-related processes (PubMed). Represented in all species,

except in prokaryotes, the function of the RING finger domain proteins is very well

conserved. The roles of ubiquitination are multiple and are essential in cellular

homeostasis, cell cycle progression and in cancer. E3 ubiquitin ligases regulate the

expression of many cell-cycle regulators which are targeted for degradation through

the proteasome pathway (Yew, 2001). Some of these cell cycle regulators, which are

over-expressed in several tumours, include cyclin D1 and E1, p27, p21, and INK4A.

Furthermore, ubiquitination plays an important role in tumorigenesis, because the

RING finger domain is found in many tumour suppressors, such as BRCA1, BARD1,

MDM2, Cbl, and Rbx1 (Freemont, 2000; Ohta and Fukuda, 2004)68. Moreover,

several nuclear oncoproteins like c-myc, c-fos, and E1A are degraded by the

ubiquitin system (Ciechanover et al., 1991). Therefore, defects in ubiquitination affect

a range of cellular processes such as cell-cycle progression, cell differentiation,

apoptosis, DNA damage pathways, and transcriptional regulation (Weissman, 2001;

Pickart, 2001; Hicke, 2001)69-71.

Ubiquitination and de-ubiquitination

Ubiquitination is a post–translational modification of proteins by attachment of

ubiquitin side chains. It is a function that modulates protein half-life by degradation of

substrates targeted with ubiquitin. Ubiquitinated proteins may either undergo

proteolysis through the proteasomal pathway or be reused after cleavage of the

ubiquitin chains by cysteine proteases, called de-ubiquitinating enzymes (DUBs). The

ubiquitin-proteasome machinery is classically composed of five essential elements:

the ubiquitin polypeptide of 76 amino acids (Ub), a ubiquitin-activating enzyme (E1),

a ubiquitin-conjugating enzyme or transferase (E2 or Ubc), a ubiquitin ligase (E3),

Page 36: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

27

and the proteasome unit (for an extensive review see Hershko and Ciechanover,

1998; short review see Wilkinson, 2000; or Freemont, 2000)53,72,73. The E1 ubiquitin-

activating enzyme binds ubiquitin by forming a thiolester bond, in an ATP dependent

way, and then transfers ubiquitin to the E2 conjugating enzyme which, in turn, forms

an isopeptide bond with the substrate via the E3 ligase. The substrates may be

ubiquitinated in several ways and therefore may serve various functions. Substrates

may be either mono- or poly-ubiquitinated (or even multiubiquitinated), and may be

linked to each other by several different residues (Ohta and Fukuda, 2004)68.

Proteins tagged with a polyubiquitin chain attached by their K48 residues are usually

targeted for degradation through the 26S proteasome pathway. However, other

polyubiquitin chains can also be synthesized with links between the C-terminus of

ubiquitin and the K6, K11, K29 and K63 on the adjacent ubiquitin. Proteins with

polyubiquitinated chains linked by their K63 residues are a signal for endocytosis, IκB

kinase activation, ribosome activation and DNA repair, whereas monoubiquitination

plays a role in the regulation of histones (Pickart, 2000)70.

BRCA1/BARD1 heterodimer is an E3 ubiquitin ligase

There are two main groups of E3 ligases, the HECT domain ligases such as the E6-

AP (Huibregste et al., 1995)74 and the RING finger ligases, such as BRCA1 and

BARD1 (Joazeiro and Weissman, 2000)75. BRCA1 is a bona fide E3 ubiquitin ligase

that may ubiquitinate its substrates or itself (Lorick et al., 1999). Cancer predisposing

mutations, C61G and C64G, within the RING domain of BRCA1, abolish the E3

ubiquitin-ligase activity of the BRCA1/BARD1 heterodimer and its tumour suppressor

function (Hashizume et al., 2001; Ruffner et al., 2001; Brzovic et al, 2001b)58,76,77.

Furthermore, these cancer mutations display hypersensitivity to γ-irradiation, showing

a link between ubiquitination, DNA repair and tumour suppression (Ruffner et al.,

2001). The stability of BRCA1 is increased in the presence of BARD1. Expression of

both is required proteins equally for high Ub-ligase activity (Hashizume et al.,

2001)76. Although, BARD1 and BRCA1 may individually carry out ubiquitination in

vitro, BRCA1 ubiquitin ligase activity is significantly enhanced when bound to BARD1

through their RING domains (Hashizume et al., 2001; Xia et al., 2003)76,78. The

aforementioned cancer predisposing mutations such as C61G and C64G perturb

binding of the E2/UbcH5c and E2/UbcH7 conjugating enzymes to BRCA1 (Brzovic et

al., 2003)59 (Figure 6). In vitro experiments with purified RING structures from

Page 37: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

28

BARD1BRCA1

Ubc5Hc binding site

A B

Ubc7H binding site

BARD1BRCA1

Ubc5Hc binding site

A B

Ubc7H binding site

unrelated proteins such as promyelocytic leukaemia protein (PML), KAP-1/TIF1B, Z,

Mel18, BRCA1 and BARD1 self-assemble in supra-molecular structures that

resemble the ones they form in cells (Kentsis et al., 2002). The same report shows

that these structures may enhance ubiquitination reactions performed by the

BRCA1/BARD1 E3 ligase and its auto-ubiquitination (Kentsis et al., 2002)79.

Therefore, the main biochemical function attributed to BARD1 is its ubiquitin ligase

activity, which is essentially mediated through its RING domain together with BRCA1.

This ubiquitination function is therefore driven mostly by the BRCA1/BARD1

heterodimer and could be responsible for mediating all cellular functions attributed to

BRCA1 and BARD1 (reviewed in Baer and Ludwig, 2002).

Ubiquitin targets of the BARD1/BRCA1 heterodimer

In the search for novel targets of the BRCA1/BARD1 E3 ubiquitin ligase function, it

was found that BRCA1 and BARD1 can be auto-ubiquitinated in vitro and in vivo in

the presence of the E2 conjugating enzyme Ubc5c and ubiquitin (Chen et al.,

2002)80. This auto-ubiquitination occurs on non-K48 residues, since a missense

mutation on this residue did not hinder poly-ubiquitination reactions in vitro. However,

BARD1 and BRCA1 auto-ubiquitination was postulated not to be a signal for

proteasomal degradation, but could rather serve other signalling pathways.

Figure 6. Representation of E2 ubiquitin conjugating binding site on BRCA1. A) Ribbon structure showing

mutation sites that perturb binding of E2 conjugating enzymes UbcH5c and UbcH7. B) Surface structure

showing that the ubiquitin conjugating enzymes only contact BRCA1 and not BARD1 (Brzovic et al., 2003).

Page 38: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

29

Furthermore, this auto-ubiquitination increases ubiquitin ligase activity of the

BRCA1/BARD1 heterodimer on itself 20-fold (Mallery et al., 2002). The auto-

ubiquitination reactions promote formation of poly-linked ubiquitin chains attached at

K6 (Wu-Baer et al., 2003; Nishikawa et al., 2003)81,82. Specifically, BRCA1/BARD1

induces formation of K6-linked conjugated ubiquitin structures at sites of DNA

replication and repair (Morris et al., 2004)83. In vitro ubiquitination reactions may be

performed with a truncated BRCA1 (1-639) (Chen et al., 2002)80 and full length forms

of BRCA1 and BARD1 (Mallery et al., 2002)84. Although the minimal interacting

regions for RING formation are BRCA1aa1-109 and BARD1aa26-119, they are not

sufficient for E3 ligase activity inferring that a larger domain is required for proper

ligase activity (Mallery et al., 2002)84. BARD1 may serve an accessory role in

ubiquitination, since the BARD1 protein or a truncated form from residues 8 to 142, in

which the RING finger domain lacked Ub-ligase activity in vitro, significantly

enhanced the E3 ligase activity of BRCA1/BARD1 duplex (Xia et al., 2003)78.

Furthermore, through NMR spectroscopic analysis and site-directed mutagenesis the

binding site of the E2 conjugating enzyme, UbcH5c or UbcH7, maps to the BRCA1

RING domain and not to BARD1 (Figure 6) (Brzovic et al., 2003)59. Although it was

shown that it is the heterodimer which is auto-ubiquitinated, no studies have shown

where and on which molecule this ubiquitination takes place and whether BARD1 is

mono-ubiquitinated. Since expression of BARD1 and BRCA1 proteins are cell-cycle

dependent and it was proposed that BARD1 and BRCA1 are degraded by the

proteasome pathway (Choudhury et al., 2004)85.

Some of the first identified targets of BRCA1/BARD1 ubiquitin ligase function are the

histones H2A and H2AX (Ruffner et al., 1998). The BARD1/BRCA1 heterodimer is

capable of catalysing the mono-ubiquitination of the histone H2A/H2AX in vitro (Chen

at al., 2002)80. Attachment of a single ubiquitin to histones H2A and H2B leads to

alteration of chromatin structure and opens DNA for transcriptional activity (Levinger

and Varshavsky 1982; Davie et al., 1990, 1991)86-88, underpinning a role for

BRCA1/BARD1 in transcriptional activation.

Another potential target for BRCA1/BARD1 ubiquitination is the FANCD2 protein,

since its mono-ubiquitination is reduced in cells that are defective for BRCA1 (Garcia-

Higuera et al., 2001)89. FANCD2 can be mono-ubiquitinated in vitro by

Page 39: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

30

BRCA1/BARD1 heteroduplex, together with E1 and UbcH5a (Vanderberg et al.,

2003)90. However, although the depletion of BRCA1 by siRNA in human cells

resulted in defective localization of FANCD2 to sites of DNA damage, it did not lead

to loss of mono-ubiquitination of FANCD2 (Vanderberg et al., 2003)90. Therefore, the

BRCA1/BARD1 E3 ligase is not necessary for mono-ubiquitination of FANCD2.

Recently some light was shed on the role of BARD1 and BRCA1 in cell-cycle events

through ubiquitination processes. Indeed, it was shown that BRCA1/BARD1 mono-

ubiquitinates γ-tubulin which regulates centrosome number and duplication cycle

(Starita et al., 2004)91. Another purported target for BRCA1/BARD1 E3 ligase is

nucleophosmin/B23 which may be ubiquitinated for its stabilization and not for

proteasomal degradation (Sato et al., 2004)92. Another a study shows that

BRCA1/BARD1 ubiquitin ligase activity is down-regulated in a cell-cycle dependent

manner by CDK2-cyclin E1/A1, but not by CDK1-cyclin B1 (Hayami et al., 2005)93. A

role for BRCA1/BARD1 ubiquitin ligase has been unveiled in DNA repair, since a

Figure 7. Ubiquitination function of BRCA1/BARD1 showing known targets of the BRCA1/BARD1 E3

ubiquitin ligase. BRCA1/BARD1 autoubiquitinates which increases ubiquitination function. The BRCC36-

BRCC45 appears also to increase autoubiquitination function.

BARD1Ub BRCA1

BRCA1/BARD1

BRCC36 – BRCC45

UbUb

UbUb

Ub

γ-tubulinUb

Nucleophosmin/B23

Ub

Rbp1

P PUb

Centrosome duplication cycle

H2A, H2AX

Ub

Transcription regulation

FANCD2

Ub

Autoubiquitination

Holo-pol II

UbChromatin decondensation

DNA repair

Cell cycle

BARD1Ub BRCA1

BRCA1/BARD1

BRCC36 – BRCC45

BRCA1/BARD1BRCA1/BARD1

BRCC36 – BRCC45BRCC36 – BRCC45

UbUb

UbUb

Ub

γ-tubulinUb

γ-tubulinUb

Nucleophosmin/B23

Ub

Nucleophosmin/B23

Ub

Rbp1

P PUb

Rbp1

P PUb

Centrosome duplication cycle

H2A, H2AX

Ub

H2A, H2AX

Ub

Transcription regulation

FANCD2

Ub

FANCD2

Ub

Autoubiquitination

Holo-pol IIHolo-pol II

UbChromatin decondensation

DNA repair

Cell cycle

Page 40: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

31

recombinant four-subunit complex comprising BRCA1/BARD1/BRCC45/BRCC36

enhanced E3 ligase activity compared to that of the BRCA1/BARD1 heterodimer

alone (Dong et al., 2003)94. These studies expand further the range of

BRCA1/BARD1 ubiquitination towards various cell cycle actions, and increase the

weight of a tumour suppressor role for BARD1 and BRCA1.

In the search for BRCA1 binding partners, BAP1, a ubiquitin hydrolase was

discovered in a yeast two-hybrid screen and interacts with the RING domain of

BRCA1 (Jensen et al., 1998)95. However, it is not clear if BAP1 competes with

BARD1 for the binding sites on BRCA1 or if it binds to the heterodimer forming a

novel functional complex. In any case, it does not appear that BAP1 is capable of de-

ubiquitinating poly-ubiquitinated BRCA1/BARD1 heterodimers in vitro (Mallery et al.,

2002). However, BAP1 could possibly deubiquitinate some of the unknown

BRCA1/BARD1 E3 targets.

IV. BARD1 functions and interacting proteins

Most breast and ovarian molecular cancer research was focused on BRCA1 and

BRCA2, as testified by more than 6000 publications on the matter during the last ten

years, compared to the 100 odd articles on BARD1. This discrepancy does not reflect

the importance that BARD1 may have in tumour suppression in gynaecological or

other tumours. However, BARD1’s function is tightly related to BRCA1 function since

it is the heterodimer which is the main functional form. However, a high number of

proteins interact or associate with BARD1 (Figure 8). BARD1 may, through BRCA1

or independently regulate, chaperone and serve as a scaffold for numerous proteins

involved in a number of cellular pathways ranging from DNA repair, transcription-

coupled repair, transcriptional regulation, apoptosis, genomic integrity and mitotic

events like centrosome duplication and cell division (Jasin, 2002)43 (Figure 9). It is

likely that most of these functions are mediated through the ubiquitin function of the

BRCA1/BARD1 heterodimer.

Page 41: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

32

Knockout mouse models and genomic instability phenotype

BARD1 shares a common feature with tumour suppressors BRCA1 and BRCA2 in

that knockout mouse follows in embryonic death and displays genomic instability

(Gowen et al.,1996; Hakem 1996; Liu 1996; Ludwig 1997; McCarthy 2003)22,23,96-98.

Cell lines lacking wild type BRCA1 are also deficient in DNA repair and homologous

recombination (Gowen et al., 1998; Moynahan et al., 1999; Scully et al., 1999)99-101.

Human BARD1 and BRCA1 counterparts of Xenopus laevis show that essential

functions as well as structure are conserved since knockouts of these genes are non-

viable (Joukov et al., 2001)49. The BARD1-null mouse embryos died in utero between

day E7.5 and E9.5 post-fecundation (McCarthy et al., 2003)98. Death was due to

severe cell proliferation defects and not to apoptosis. Partial rescue was obtained in

BARD1-/-;p53-/- double knockout embryos, since embryonic death was delayed until

day E9.5. Cytogenetic analysis of the BARD1-/-; p53-/- cells displayed an increase of

structural and numerical chromosome aberrations compared to p53-/- cells

(McCarthy et al., 2003)98. Furthermore, conditional knockout experiments of BARD1

results in mammary cancers, chromosomal instability and proliferation defects

(personal communication Thomas Ludwig) and manifest the same characteristics of

human breast cancers. The drastic phenotypes of the knockout mice means that

BARD1 is essential for a wide panel of cellular events (reviewed by Jasin, 2002)43. It

is clear that BARD1 assures vital functions for cell cycle progression and

maintenance of genome integrity (Figure 9).

Figure 8. Illustration showing the known proteins that interact with BARD1. Proteins that are thought to

have a direct interaction are shown in colour. Proteins that are known to bind indirectly with BARD1 are

shown in grey, all of which may be detected in biochemical complexes.

p53

BRCA1Bcl-3

RAD 51NF-κB/p50

CstF-50

hMSH2-hMSH6 BRCC36-BRCC45

EWS

BASC

RING BRCTANKYRIN

MRE11

RAD50

NBS1

Holo-pol II

p300/CBP

BRCA2

Ku-70

TACC1

AuroraB

XIST

IκBα

p53

BRCA1Bcl-3

RAD 51NF-κB/p50

CstF-50

hMSH2-hMSH6 BRCC36-BRCC45

EWS

BASC

RING BRCTANKYRIN

MRE11

RAD50

NBS1

Holo-pol II

p300/CBP

BRCA2

Ku-70Ku-70

TACC1

AuroraB

XIST

IκBαIκBα

Page 42: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

33

BARD1 in DNA-damage repair

It has been established that BARD1, in conjunction with BRCA1, is essential for

maintaining chromosomal stability. This is done partly through mechanisms of

homology-directed repair of DNA double-strand breaks. In fact, BRCA1 regulates a

variety of DNA damage pathways. Cells lacking functional BRCA1 have severe

defects in transcription-coupled repair, homology-directed repair (HDR), non-

homologous-end-joining (NHEJ) and microhomology-end-joining repair pathways

(Gowen et al., 1998; Abbott et al., 1999; Moynahan et al., 1999; Snouwaert et al.,

1999; Zhong et al., 2002a, 2002b)99,100,102-105. It is now clear that BARD1 is also

reqired in those DNA repair pathways (Westermark et al., 2003)106. It is the

BRCA1/BARD1 heterodimer which is the functional unit in DSB repair (Jasin,

2002)43. However, it is not known whether BARD1 plays other roles in these

pathways.

A functional interaction between hMSH2 and BRCA1 was discovered in a yeast-two-

hybrid screen, which was confirmed by co-localization and immunoprecipitation

studies (Wang et al., 2001)107. Further investigation showed that hMSH2 associates

with BARD1. In fact, BARD1 may also associate with other enzymes of the same

pathway such hMSH3 and hMSH6. The interaction of BARD1 with hMSH2-hMSH6

substantiates a role for BARD1 and BRCA1 in DNA mismatch repair, and suggests

that BARD1 cannot be dissociated from its partner in DNA damage sensing and

signalling (Wang et al., 2001)107. It is noteworthy that BRCA1 interacts with numerous

other proteins involved in the repair pathways, the list is long, but one may note that

ATM, ATR, DNA-PKs, Ku-70 and Ku-80 are involved in sensing DNA breaks and

signalling through phosphorylation of BRCA1 that the DNA repair pathways should

be switched on. Lastly, it would be interesting to evaluate the participation of BARD1

in all these interactions. Finally, it was shown that the BRCA1/BARD1 orthologues in

C. elegans are also required for DNA repair (Boulton et al., 2004)50, supporting the

thesis that these proteins have highly conserved functions in different species.

Cellular localization of BARD1 and BRCA1 during DNA repair

At the cellular level, the expression levels of BARD1 protein were first reported to be

relatively constant throughout the cell cycle, in contrast to the elevated levels of

Page 43: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

34

BRCA1 in S-phase (Jin et al. 1997)108. Subsequently, it was shown that cells that

were arrested in G0-G1 by serum withdrawal express low levels of BRCA1. The

levels progressively accumulate after release from starvation in G1 phase, peaking at

the G1/S border. They remain high during S-phase, mirroring cyclin A expression

levels (Chen et al., 1996; Vaughn, 1996; Gudas et al., 1996; Rajan et al. 1996; Jin et

al., 1997)108-112. These reports obviously suggest a role for BRCA1 during S phase.

However, further studies show that BARD1 expression levels fluctuate in a cell-cycle-

dependent manner, with an increase during mitosis (Jefford et al., 2004; Choudhury

et al., 2004; Hayami et., 2005)51,85,93. Concomitant expression of BARD1 and BRCA1

was observed during S-phase (Hayami et al., 2005)93. Evidence of a BRCA1/BARD1

interaction came from immunofluorescence experiments, where BARD1 co-localized

with BRCA1 and Rad51 during S-phase in “nuclear foci” (Scully et al., 1997b, Jin et

al., 1997 108,113. Upon DNA damage by hydroxyurea (HU), the BRCA1 and BARD1

nuclear dots disperse, but re-aggregate at new foci containing PCNA (Proliferating

Cell Nuclear Antigen). The BRCA1/BARD1/Rad51 appears to relocate to areas of

DNA damage and play a role in repairing damaged DNA (Scully et al., 1997b)113.

These data are supported by other findings, demonstrating that BRCA1/BARD1

heteroduplex associates with DNA at stalled replication forks (Paull et al., 2001)114.

The loss of the BRCA1 foci was accompanied by a dose-dependent increase of

BCRA1 phosphorylation. Concordantly with these findings, BARD1 was

biochemically identified with BRCA1 in hydroxyurea-inducing complexes (HUIC)

formed after treatment with HU (Chiba and Parvin, 2001)115. The BRCA1 protein was

reported to interact with the Mre11-Rad-50-Nbs1 (MRN) repair complex (Zhong et al.,

1999)116. However, the HUIC complex is distinct from the BRCA1-Mre11-Rad50-

Nbs1 complex (Chiba and Parvin, 2001)115.

BARD1 in transcriptional regulation and chromatin remodelling

Because BRCA1 and BARD1 share a homologous BRCT domain, the first clues of a

purported role for BARD1 in transcription derived from transient transfection

experiments with BRCA1. BRCA1 was shown to be able to activate transcription in

vitro (Chapman et al., 1996; Monteiro et al., 1996, 1997)117-119. Further evidence

showed that the BRCT domains of BRCA1 induces transcriptional activation in vitro

(Haile et al., 1999)120. Additionally, BRCA1 interacts with the RNA polymerase II

holoenzyme (holo-pol) in association with RNA Helicase A, a protein also involved in

Page 44: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

35

transcription (Anderson et al., 1998)121. On the other hand, it is not known whether

BARD1 is required for these transcription-related interactions. In later experiments, it

was shown that BRCA1 can bind to DNA at sites of DNA branching (Paull et al.,

2001)114. This direct binding is another clue for BRCA1-driven transcription. However,

conclusive evidence for a BRCA1-induced transcription activation came from

biochemical experiments demonstrating that BRCA1 is a bona fide component of the

RNA polymerase II holoenzyme, since BRCA1 co-purifies with the holoenzyme

complex (Scully et al., 1997; Neish et al., 1998)122,123. Further studies showed that

BARD1 was also part of this multimeric complex (Chiba and Parvin, 2002)124,

implicating BARD1 in transcriptional regulation. By using deletion mutants of BRCA1,

it was also shown that the N-terminal end of BRCA1 containing the RING domain

was necessary for the activity of the holoenzyme, whereas the C-terminal end

deletion showed little reduction in regulation of transcription (Chiba and Parvin,

2002)124. This latter fact also implies that BARD1 is the linking protein tethering

BRCA1 to the holo-pol thereby confirming BARD1’s role as transcriptional co-factor.

More evidence on BARD1’s role in transcription came from the discovery that BARD1

interacts with the NF-κB/Rel transcription factors. The NF-κB signalling pathway is

conserved through evolution. The NF-κB transcription factors play numerous, crucial

roles in inflammation, immunity and apoptosis; being over expressed in many

tumours (Baldwin 1996, 2001)125,126. BARD1 binds to Bcl-3, which is part of the IκB

family of NF-κB inhibitors. Bcl-3 serves as a bridge between the p50, one of the five

known components of the NF-κB co-factors, and BARD1 (Dechend et al., 1999)127.

This binding could be mediated through Bcl-3’s and BARD1’s ankyrin domains.

Furthermore, it has also been shown that the p65/RelA subunit of NF-κB binds to

BRCA1, and so causing an increase of the transcription levels of NF-κB target genes,

such as Fas and interferon β (IFNβ) (Benezra et al., 2003)128. Transcription can be

diminished by BARD1 and a BRCA1 RING domain mutants (Benezra et al., 2003)128.

The plot between NF-κB and BARD1 thickens, because BARD1 has a slight binding

affinity with IκBα (Dechend et al., 2003)127 and because IκBα is ubiquitinated for

degradation by the proteasome after being phosphorylated on serines 32 and 36.

(reviewed in Chen and Greene, 2003)129. The question remains open as to whether it

is the BARD1/BRCA1 E3 ubiquitin ligase that is responsible for tagging IκBα with

Page 45: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

36

ubiquitin. However, the complexity of these pathways, (which are never linear),

increases, because p50 and RelA are acetylated at multiple serine residues probably

by the CBP/p300 acetyltransferase, which was evidenced by BRCA1 and CBP/p300

interactions seen in vitro and in vivo (Cui et al., 1998; Pao et al., 2000)130,131. It is also

noteworthy that CBP (CREB binding protein) is a component of the polymerase II

holoenzyme (von Mikecz et al., 2000)132. BARD1 is the main co-partner of BRCA1 in

a CBP/p300 mediated transcription. The cAMP response element binding protein

(CREB) binding protein (CBP) and p300 are paralogous proteins involved in

transcriptional regulation because they may bind CREB and other transcription

factors like TFIIB and TATA box-binding protein. The CBP/p300 can function as a co-

activator in a BRCA1-mediated transactivation (Pao et al., 2000)131. The CBP/p300

proteins are also histone acetyltransferases (HATs) (Ogryzko et al., 1996; Bannister

and Kouzarides, 1996)133,134, and may recruit other HAT proteins. Acetylation of N-

terminal lysine residues of histones enables chromatin decondensation so favouring

transcriptional activation. Other factors are also involved in chromatin remodelling.

The SWI/SNF-related complex binds to BRCA1 further confirming a role for

BRCA1/BARD1 in transcriptional activity (Neish et al., 1998; Bochar et al.,

2000)123,135. In the same line of thought, BARD1 and BRCA1 may also be directly

involved in heterochromatin re-structuring and in X chromosome inactivation because

BRCA1 and BARD1 co-associate with XIST RNA (Ganesan et al., 2002)136. It was

first discovered that BRCA1 decorates the unpaired X chromosome in human

pachytene spermatocytes upon staining of meiotic chromosomes (Scully et al.,

1997b)137.

BARD1 in mRNA processing

BARD1, together with BRCA1 interacts with the mRNA polyadenylation factor CstF-

50 and represses the activity of the polyadenylation machinery in vitro (Kleiman and

Manley, 1999)138. The cleavage stimulation factor (CstF) is a protein complex

involved in the polyadenylation process of all eukaryotic mRNAs. The CstF is

required for the endonucleolytic cleavage step of mRNA and it helps to properly

identify the site of processing (Takagaki, 1990,1997)139,140. It is a heterotrimeric

protein with subunits of 77, 64 and 50 kDa. In a yeast two hybrid screen using CstF-

50 as bait, BARD1 was identified as having one of the strongest interactions with

CstF-50 (Kleiman and Manley, 1999)138. Furthermore, CstF/BARD1/BRCA1 complex

Page 46: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

37

Figure 9. The BARD1 network. This illustration shows the various pathways in which BARD1 is involved.

BARD1 is up-regulated upon DNA damage, genotoxic insult and in the G2/M phase of the cell cycle.

BARD1 down-regulation induces a genetic instability phenotype and a tumour phenotype. BARD1 is

involved in DNA repair pathways, in replication, and in transcriptional regulation.

P

BARD1 P

BRCA1

BARD1

CDK1/2

UUb

Increase of BARD1 in G2/MBARD1

ApoptosisBARD1

p67 cleavage product

Inhibition of polyadenylation

Relocation in PCNA aggregates

p53

Genetic Instability Tumour phenotype

DNA repair, MMR, NHEJ, HDR

MRE11

RAD50

NBS1

RAD51

BASC

BARD1 repression

HU-UV

Holo-pol II

Replication S-phase

Transcriptional block

RAD51 PCNA

CstF-50

Transcription regulation

Bcl-3

NF-κB/p50

DNA

Ub

ATM-ATR-DNA-PKs DNA damage

hMSH2-hMSH6

XIST X silencing

BRCA2-TACC1-

Aurora B Cytokinesis

calpain

is involved in the inhibition of mRNA 3’ processing in cell lines that were treated with

DNA damage-inducing agents such as HU or UV (Kleiman and Manley, 2001)141.

Furthermore, addition of exogenous BARD1 may also inhibit polyadenylation.

Regulation of polyadenylation could play an important role in the control of cellular

proliferation (Colgan et al., 1997; Zhao and Manley, 1998)142,143. It is therefore

possible that BARD1-mediated inhibition of CstF polyadenylation prevents specific

RNA processing during DNA damage-induced repair, thus helping to stall the cell

cycle during a BRCA1/BARD1-mediated DNA repair. Further evidence of the

involvement of BARD1 in tumour suppression through mRNA processing, came from

the fact that the Ewing’s sarcoma gene product EWS interact via its NH2 terminus to

BARD1 (Spahn et al., 2002)144. However, its modus operandi remains speculative.

Page 47: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

38

BRCA1-independent pro-apoptotic functions of BARD1

Since BARD1 and BRCA1 were not consistently expressed in all tissues, differential

expression led to the hypothesis that BARD1 might play a role in a BRCA1-

independent function (Irminger-Finger et al., 1998)47. Indeed, the first clues on the

crucial role of BARD1 function in tumorigenesis, independently of BRCA1, came from

BARD1 repression experiments, through ribozyme and mRNA anti-sense production

in a mouse mammary epithelial cell line in vitro culture. The abrogation of BARD1

expression induced a spectrum of biological alterations, such as drastic

morphological cell change, loss of contact inhibition, aberrant cell-cycle progression

and abnormal DNA content, suggestive of a premalignant phenotype and genomic

instability(Irminger-Finger et al., 1998)47. These data were later confirmed by others

showing chromosomal instability in BARD1-/- cells as evidenced by aneuploidy and

abnormal karyotypes (Joukov et al., 2001; McCarthy et al., 2003)49,98. Indeed, the

main characteristic of cells with a deficient BRCA1 or BARD1 expression, following

RNA interference treatment, is an increase of chromosomal alterations, as observed

by cytogenetic analysis of karyotypes. Depletion of BRCA1 mRNA by siRNA

expression can cause genomic instability in Fanconi anaemia cells (Bruun et al,

2003)145.

Since BARD1 has a vital role in normal cell survival, BARD1 expression was

measured during apoptosis. It was found that BARD1 expression, through

transcription, was up-regulated in response to genotoxic stress or after induction of

an ischemic stroke in mouse (Irminger-Finger et al. 2001)146. Additionally, over-

expression of exogenous BARD1 leads to DNA fragmentation and caspase-3

activation indicative of apoptosis (Irminger-Finger et al., 2001)146. It is dependent of

p53, but independent of BRCA1, since BARD1 may trigger apoptosis in the absence

of BRCA1 and not in p53 deficient cell lines (Irminger-Finger et al. 2001)146. These

experiments identified BARD1 as a mediator between pro-apoptotic stress and p53-

dependent apoptosis. The fact that Q564H missense mutation of BARD1 was

defective in inducing apoptosis when transfected in cultured cells, implies that the

region of BARD1 around Q564H could serve a role necessary for tumour

suppression and for the pro-apoptotic function of BARD1. Furthermore, it was also

shown that BARD1 co-immunoprecipitates with p53, suggesting a physical interaction

Page 48: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

39

between p53 and BARD1 in order to induce apoptosis. BARD1 up-regulation upon

genotoxic stress is accompanied by p53 expression and stabilization (Irminger-Finger

et al., 2001). Activation and stabilization of p53 depend on its phosphorylation at

multiple sites, by a number of kinases, including phosphorylation on serine 15.

(Meek, 1994; Milczarek et al., 1997)147,148. Moreover, the stabilization and

phosphorylation of p53 may be performed through BARD1, since the absence of

functional BARD1 is sufficient for abolishing p53 phosphorylation on serine 15

(Fabbro et al, 2004)149. It was shown that the BRCA1/BARD1 heterodimer is required

for phosphorylation of p53 on serine 15. This phosphorylation of p53 is necessary for

a G1/S cell cycle arrest following DNA damage induced by ionising radiation (Fabbro

et al., 2004a)149. Further evidence showed that BARD1 overexpression can catalyse

the phosphorylation of p53 on serine 15 by a DNA-damage response kinase (Feki et

al., 2005)150. It was also shown that BARD1 binds to Ku-70 in co-immunoprecipitation

experiments as well as phosphorylated p53, identifying a candidate kinase for p53

phosphorylation (Feki et al., 2005)150. Identification of the region of BARD1 which

binds to p53 was performed through a series of co-immunoprecipitations experiments

of deletion mutants fused to EGFP. BARD1 may co-immunoprecipitate p53 over a

large, domain but not the RING finger domain. Even the deletion mutants lacking the

RING functional domain, which normally binds BARD1 to BRCA1, are capable of co-

immunoprecipitating p53 (Feki et al., 2005)150. These experiments exclude the

preconception that p53 was pulled down through an interaction with BRCA1. The p53

protein must be phosphorylated on serine residues for a function in apoptosis. It

appears that BARD1 stabilizes p53 for DNA repair and for apoptosis.

BRCA1 may also bind to p53 in two regions: to the region 224-500 and to the BRCT

domain 1760-1863 (Chai et al., 1999)151. This interaction is functional since BRCA1

may regulate p53-dependent gene expression (Ouchi et al., 1998)152 and that,

conversely, p53 is required to modulate BRCA1 expression (Arizti et al., 2000)153. It

appears that BARD1 follows a similar BRCA1 behaviour since BRCA1 over-

expression triggers apoptosis (Harkin et al., 1999)154 and causes growth retardation

in several breast or ovarian cell lines (Holt et al., 1996; Jensen et al., 1996;

1998)21,95,155.

Page 49: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

40

Mapping of the pro-apoptotic domain of BARD1

BARD1’s function in the nucleus is mainly linked to DNA repair and transcriptional

activity, but little is known on its dynamic properties and its role within the cytoplasmic

compartment. BARD1, like BRCA1, has been described until now mostly as nuclear

protein. However, BARD1 is synthesized in the cytoplasm and translocated into the

nucleus localizing to foci (Scully et al., 1997)113. Human BARD1 has 6 predicted NLS

(Nuclear Localizing Signal) sequences, whereas the mouse has only 3 NLSs. In both

species, the prediction score for nuclear localization is high. BARD1 even serves as a

chaperone for BRCA1, by retaining BRCA1 in the nucleus by masking BRCA1’s NES

(Nuclear Export Signal) sequence when the two are bound together (Rodriguez et al.,

2000; Fabbro et al., 2002; Schuechner et al., 2005)156-158.

However, contrary to previous reports stipulating that BARD1 had an exclusive

nuclear localization, we showed that BARD1 is also localized in the cytoplasm in vitro

and in vivo, albeit in discreet levels (Jefford et al., 2004)51. We showed that apoptosis

is concomitant with enhanced cytoplasmic localization of BARD1, as well as the

appearance of a p67 cleavage product (Jefford et al., 2004)51. Moreover similar

observations suggest that BARD1 can shuttle from the nucleus to the cytoplasm,

increasing its apoptotic activity (Rodriguez et al.,2004)159. Mapping the apoptotic

domain of BARD1 was performed by using deletion constructs, which confirmed the

apoptotic-inducing region to residues 510-604 in between the ankyrin repeats and

the BRCT domains. This region coincides with two known cancer predisposing

mutations C557S and Q564H (Thai et al., 1998; Ghimenti et al., 2002; Karppinen et

al., 2004; Jefford et al., 2004)45,51,160,161. This region is necessary for BARD1-induced

apoptosis, but it is not known if it is sufficient. BARD1 post-translational modifications

and its degradation have not been investigated thoroughly, but it was shown that a

p67 BARD1 proteolytic product cleaved by calpain can act as a tumour antigen. It is

associated with apoptotic bodies in rat colon cancer cells (Gautier et al., 2000)162.

BRCA1 and BARD1 functions in cell cycle

BARD1 and BRCA1 have several roles in cell-cycle progression that are mediated

through their ubiquitin function and controlled by a cyclin dependent kinase, CDK2

(Hayami et al. 2005). Furthermore, BARD1 is phosphorylated by CDK1/2 on its NH2

terminus (Hayami et al., 2005)93. A role for BRCA1 in centrosome duplication has

Page 50: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

41

been described, since BRCA1 associates with centrosomes during mitosis (Hsu and

White, 1998)39 and co-immunoprecipitates with γ-tubulin. The γ-tubulin binding site

was mapped to exon 11 of BRCA1 (Hsu and White, 2001)40. BRCA2 is also

implicated in stages of cytokinesis, since depletion of BRCA2 increases cell division

defects (Daniels et al., 2004)163. To further ascertain the role of BARD1 in cell cycle

progression and cytokinesis, immunoprecipitation studies were performed in our

laboratory and show that BARD1 co-aggregates withTacc1/AuroraB/BRCA2 at the

midbody during completion of cytokinesis (Jefford and Irminger-Finger, submitted).

Furthermore, this study has also shown that depletion of BARD1 clearly disrupts cell-

cycle progression, cytokinesis and creates cellular fusion. These results provide an

explanation for the mechanisms leading to aneuploid karyotypes and chromosomal

instability as a cause for carcinogenesis in cells with deficient expression of BARD1

and BRCA1.

Figure 10. Mutations in BRCA1 span the entire gene. A total of 657 BRCA1 breast and ovarian cancer

predisposing mutations have been documented.

(Human Gene Mutation Database: http://archive.uwcm.ac.uk/uwcm/mg/hgmd0.html)

Page 51: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

42

V. BARD1 mutations and expression in cancer

BARD1 tumour suppressor mutations in breast and ovarian cancers

It was demonstrated that BRCA1 and BRCA2 mutations predispose women to cancer

(Miki et al., 1994)5, but do not account for all familial breast and ovary cancer

predispositions. The cancer predisposing mutations of BRCA1 span the entire gene

and are numerous (Figure 10). A specific cancer predisposing mutation, C61G, within

the RING domain of BRCA1, disrupts formation of the BRCA1 homodimer (Figure 4b)

(Brzovic et al., 1998)164. Moreover, this mutation impedes ubiquitin protein ligase

activity and tumour suppressor function of the BRCA1/BARD1 heterodimer

(Hashizume et al., 2001; Ruffner et al., 2001)76,77. It was therefore hypothesized that

BARD1 mutations should also predispose to cancer. After screening a panel of

sporadic breast, ovarian and endometrial cancers, three missense alterations were

identified in the BARD1 gene at the amino acid positions Q564H, V695L, and S761N

(Thai et al., 1998)45. Loss-of-heterozygosity was accompanied in two cases (Q564H

and S671N), substantiating BARD1’s role as a tumour suppressor. The V695L and

S761N mutations arose in somatic breast tissue and were not found in the germ-line,

whereas the Q564H mutation was found in the germ line of a patient with clear cell

adenocarcinoma of the ovary (Thai et al., 1998)45. The latter mutation was screened

against a panel of over 300 healthy individuals, suggesting that this alteration is not a

polymorphism. Five alterations were discovered in an Italian cohort with familial

breast and ovarian cancers that was chosen for its absence of BRCA1 and BRCA2

gene alterations in its proband (Ghimenti et al., 2002)160. These mutations include 3

missense mutations, K312R, C557S, N295S, and an in-frame deletion of 7 amino

acid residues, 1139Del21-(PLPECSS). The last alteration is a C1579G transversion

with no amino acid change at position A502, which was found in 15 probands,

indicative of a novel polymorphism variant (Ghimenti et al., 2002)160. The mutations

C557S and 1139Del21 were described previously by Thai et al., (1998) but were

considered as polymorphisms. However, segregation analysis showed that the

C557S mutation may be linked with the tumour with a statistically borderline

significance (Ghimenti et al., 2002)160. This mutated allele was further confirmed as

being involved in hereditary susceptibility to breast cancer in a Finnish population of

gynaecological cancers (Karppinen et al., 2004)161. Individuals carrying the C557S

allele showed a significantly increased risk of breast cancer. The observed incidence

Page 52: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

43

of 7.4% for breast cancer against 1.4% in healthy controls suggests that it could be a

common susceptibility allele with low penetrance (Karppinen et al.,2004)161.

Mutational analysis of BARD1 in familial breast cancer patients in Japan, that were

tested negative for BRCA1 or BRCA2 germ-line mutations, revealed six alterations:

(S241C, G378S, N470S, V507, 1139Del21) (Ishitobi et al., 2003)165. Of the six

mutations, only one missense mutation, N470S, was found in the germ-line of a

patient and defined as a cancer predisposition mutation, because it did not appear in

a large background of healthy controls (Ishitobi et al., 2003)165. So far, these

observations show that BARD1 germ-line mutations have a low association with non

BRCA hereditary breast and ovary tumours and account for a very small fraction of

familial breast cancers (Figure 11). This fact may question the implication of BARD1

in predisposition to gynaecological cancers. However, expression of BARD1 is

necessary in developing embryos, since homozygous Bard1 knockout mice die in

utero at an early stage (McCarthy et al., 2003)98 and that the conditional knockout of

the Bard1 allele causes tumours in mammary glands several months after induction

(personal communication Thomas Ludwig).

BARD1 expression in normal and tumour tissues

The regulation of transcription of BARD1 has not been totally elucidated, but BARD1

gene structure spans a 10 kb region close to the telomere on chromosome 2q34-q35

(Wu et al., 1996; Thai et al., 1998)44,45. BARD1 cDNA is 2530 base pairs in length

and is composed of 11 exons. Several splice variants have been identified for

BRCA1 (Orban and Olah, 2000; McEachern et al.,2003)166,167. Likewise, BARD1 has

also several transcripts supplied by alternative splicing (Feki et al., 2004, 2005)150,168.

Nevertheless, BARD1 and BRCA1 mRNA expression patterns as observed from

northern blots are coordinated in several tissues.

In most murine tissues, Bard1 and Brca1 are concomitantly expressed. Northern blot

experiments showed that BARD1 RNA messengers were abundantly expressed in

spleen and testis, but not in heart, brain, liver, lung, skeletal muscle or kidney tissues

(Ayi et al., 1998)48. RNase protection assays showed high expression levels in

spleen and testis, but that Bard1 was also expressed in various other murine tissues,

but less than Brca1 (Irminger-Finger et al., 1998)47. Hormonally regulated organs

were also assayed for mBard1 expression, and showed that Bard1 in uterus is

Page 53: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

44

increased from di-oestrus through post-oestrus phase, whereas Brca1 increases from

diestrus to early oestrus and decreases during oestrus and post-oestrus (Irminger-

Finger, 1998)47. This non-coordinate expression of Bard1 with Brca1 was the first

indication of a BRCA1-independent role of BARD1. In our laboratory, Bard1

expression in testis was further explored, showing that Bard1 is expressed at all

stages of spermatocyte maturation, whereas Brca1 expression is only seen in meiotic

and early round spermatocytes (Feki et al. 2004)169. Furthermore, it was shown that

BARD1 transcripts result from alternative splicing. The messenger RNA mBARD1β

and mBARD1γ from rat testis, could have a role in triggering apoptosis (Feki et al.,

2004)169. Finally it was also found that cells from the cytotrophoblast express also

splice variants for BARD1 and protein expression was detected in the extra cellular

matrix. Secretion of BARD1 is presently being investigated (personal communication

by Bischof and Irminger-Finger).

Expression of BRCA1 mRNA is often reduced in sporadic breast carcinomas

(Thompson et al., 1995)170, even in the absence of distinct BRCA1 mutations. This

gene silencing anomaly is often imputed to epigenetic effects such as BRCA1

promoter methylation on CpG rich islands (Mueller and Roskelley, 2003)34. Abnormal

2530 bp

5’UTR 3’UTRRING Finger ANKYRIN REPEATS BRCT DOMAINS

Putative cancer predisposing mutations (7)

Polymorphisms (7)

Silent mutations (4)

* Somatic mutations

i ii iii iv v vi vii viii ix x xi

777 aaP24S K153E V507M

Q564H S761N *C557S

Del 358-364

V695L *

S241C

N470SN295S

R378S

L312N

R658C

G203 T351 H506A502

1 (74)

LOH LOH

2530 bp

5’UTR 3’UTRRING Finger ANKYRIN REPEATS BRCT DOMAINS

Putative cancer predisposing mutations (7)

Polymorphisms (7)

Silent mutations (4)

* Somatic mutations

i ii iii iv v vi vii viii ix x xi

777 aaP24S K153E V507M

Q564H S761N *C557S

Del 358-364

V695L *

S241C

N470SN295S

R378S

L312N

R658C

G203 T351 H506A502

1 (74)

LOH LOH

Figure 11. Diagram showing all the documented mutations pertaining to BARD1. No cancer predisposing

mutations were found in the RING domain. A total of 7 mutations were found in a population of cases of breast

cancer that were negative for BRCA1 or BRCA2, which were screened against a large background of healthy

individuals. Only two cases of loss-of-heterozygosity were found in the tumours of breast cancer patients (Q564H

and S761N). (Compiled from Thai et al., 1998; Ghimenti et al., 2002; Ishitobi et al., 2003; Karppinen, 2004).

Page 54: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

45

reduction of BRCA1 expression was observed in about 30% of breast carcinomas

cases in a Japanese population (Yoshikawa et al., 2000)171. In contrast, expression

of BARD1 and of the two mismatch repair enzymes, hMLH1 and hMSH2, was not

reduced as frequently (Yoshikawa et al., 2000)171. Furthermore, p53 expression in

sporadic breast cancers is diminished, which is accompanied by the frequent

detection of a mutated p53 (Feki and Irminger-Finger, 2004)168. A recent report has

examined the expression pattern of several genes, Smad7, Smad2 and Bard1,

implicated in the TGFβ signalling pathways of normal and breast cancers. The

expression of these genes are induced by TIEG (TGF-β inducible early gene) which

is a nuclear zinc-finger transcription factor (Rheinholz et al., 2004)172. The expression

was measured by real-time RT-PCR and the result was a decrease in breast cancer,

which supports the notion that these genes are tumour suppressors repressed in

tumours (Rheinholtz et al., 2004)172. Another study also showed a diminution of

BARD1 expression at the mRNA level in breast carcinomas (Qiu et al., 2004)173.

However, this contrasts with a recent study, where it was found that a mutated form

of BARD1 was over-expressed in a number of cancerous breast tissues and its

expression was localized in the cytoplasm (Jian-Wu et al., 2005 in press). This

abnormal cytoplasmic over-expression could be due to a dysfunctional negative

feedback loop, which explains the accumulation of BARD1 in the cytoplasm, since its

down stream effectors are non-responsive. These apparently contrasting results

demonstrate the discrepancies between mRNA expression and protein expression.

The explanation could be that in abnormal or precancerous cells BARD1 up-

regulation is necessary for a signalling towards apoptosis, resulting in cytoplasmic

localization for programmed cell death signalling. However, since the tumour tissues

do not respond to the BARD1 protein, BARD1 accumulates in the cytoplasm, which

results in a negative feedback loop thereby switching off mRNA transcription.

VI. Conclusion

BARD1 activities reflect the mirror image of BRCA1 functions in DNA repair,

transcription and in maintenance of genome integrity, because BARD1 may perform

the cellular roles imputed to BRCA1. However, on the other side of the coin, BARD1

may be involved in several BRCA1-independent functions, just as BRCA1 may have

Page 55: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

46

several BARD1-independent functions. Indeed, BARD1, with differential expression,

has functions in DNA repair mediating, through p53, but independently of BRCA1, an

apoptotic response. BARD1 is possibly the effector or the co-activator in an

apoptosis-signalling pathway, since it is up-regulated in apoptosis, it contributes to

p53 phosphorylation and its over-expression may induce apoptosis.

Clearly, ordinate cellular functions, mediated by the heterodimer, must rely on the

BARD1/BRCA1 equilibrium. BRCA1 stabilizes BARD1 and vice versa. For instance,

BRCA1 over-expression induces BARD1 up-regulation. Moreover, it is worth noticing

that BARD1 post-translational expression is modified in the absence of BRCA1, since

BARD1 loses its ubquitinating functions and its ubiquitin moiety.

Like BRCA1, BARD1 interacts with a high number of proteins involved in various

cellular functions. BRCA1 and BARD1 have several ubiquitination targets that affect

many and various systems. It is likely that BRCA1 and BARD1 serve as a platform for

coordinating a large number of cellular functions, probably through its main

ubiquitinating functions. The BARD1 and BRCA1 proteins, either together or

independently, play important roles in various systems, which include DNA-related

functions (DNA repair, replication, chromatin remodelling and transcription) and

mitotic events (centrosome duplication and checkpoint kinases). Thus, BARD1 and

BRCA1 could serve as the main mediator between the DNA-related functions and

cell cycle events.

The multitude of interacting proteins and pathways that BARD1 and BRCA1 are

involved in is flabbergasting. The unifying model proposes that all the pathways that

BARD1 is involved in converge towards a unique goal, an integrated system

safeguarding the integrity of the genome. Epigenetic effects of BARD1 and BRCA1

may be partly explained from this point of view. In all cases, it looks like BARD1 and

BRCA1 are at the crossroads of multiple pathways coordinating different systems in

an integrated network.

Page 56: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

47

Results Publications

1. Identification of BARD1 as a mediator between proapoptotic stress and p53-dependent apoptosis

Irminger-Finger I, Leung WC, Li J, Dubois-Dauphin M,Harb J, Feki A, Jefford CE,

Soriano JV, Jaconi M, Montessano R, Krause KH. Mol Cell. 2001 Dec; 8(6):1255-66.

This seminal paper shows that BARD1 is a mediator between proapoptotic stress

signals and induction of programmed cell death through p53 activation. Although my

contribution was small, it is a crucial report since its premises serves as a basis of

subsequent work and constitutes the hypotheses for my thesis. In this study, we

showed that BARD1 expression at the mRNA and protein levels is augmented upon

genotoxic stress in vitro, and by hypoxia in vivo in cerebral ischemia. Upregulation of

BARD1 expression upon cell death induced by genotoxic insult, such as UV and

doxorubicin treatment, is accompanied by elevated p53 expression in mouse

mammary gland and embryonic stem cells. Cells devoid of functional BARD1 appear

not to undergo apoptosis as measured by DNA fragmentation assays and caspase-3

activity. We showed that BARD1 over-expression in transient transfection

experiments may induce apoptosis even in the absence of a functional BRCA1, but

not in the absence of p53, which identifies BARD1 as a bona fide signalling molecule

towards apoptosis. BARD1 induces apoptosis in concentration-dependent and time-

dependent manners. The pro-apoptotic activity of BARD1 coincides with p53

elevation. Furthermore, we showed that p53 co-immunoprecipitates with BARD1,

indicative of a physical interaction between BARD1 and p53. This interaction implies

the participation of BARD1 in the signalling towards cell death. My contribution

consisted in the various assays, the preparation of cells, transfection experiments,

and the repetition of the functional apoptotic assays, in which apoptosis was

measured by DNA fragmentation assays. Furthermore, I also worked on optimising

the BARD1-p53 co-immunoprecipitation assays.

Page 57: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Molecular Cell, Vol. 8, 1255–1266, December, 2001, Copyright 2001 by Cell Press

Identification of BARD1 as Mediatorbetween Proapoptotic Stressand p53-Dependent Apoptosis

nant phenotype and genomic instability (Irminger-Fingeret al., 1998). The BARD1/BRCA1 complex was shownto associate with the polyadenylation factor CstF-50(Kleiman and Manley, 1999), presumably to prevent RNAprocessing at sites of DNA repair (Kleiman and Manley,

Irmgard Irminger-Finger,1,7 Wai-Choi Leung,5

Jian Li,1 Michel Dubois-Dauphin,2 Jean Harb,6

Anis Feki,1,3 Charles Edward Jefford,1

Jesus V. Soriano,4 Marisa Jaconi,1

Roberto Montesano,4 and Karl-Heinz Krause1

1 Biology of Aging Laboratory 2001). The BRCA1-BARD1 heterodimeric RING fingercomplex was recently shown to possess ubiquitin ligaseDepartment of Geriatrics

2 Department of Psychiatry activity in vitro (Hashizume et al., 2001). However,BARD1 and BRCA1 are not consistently coexpressed3 Department of Gynecology and Obstetrics

4 Department of Morphology in different tissues (Irminger-Finger et al., 1998), sugges-tive of the existence of independent functions for bothFaculty of Medicine

University of Geneva genes. Of particular interest in this context is the pre-viously observed high expression level of BARD1 butSwitzerland

5 Department of Pathology and Tulane Cancer Center undetectable BRCA1 concentrations in mouse uterusduring the estrus phase (Irminger-Finger et al., 1998),Tulane University School of Medicine

New Orleans, Louisiana i.e., the moment of the estrous cycle when massive celldeath (in part through apoptosis) of the endometrium6 Institut de Biologie

INSERM U419 occurs (Ferenczy, 1993). As induction of apoptosis isan important mechanism for tumor suppression, we hy-Nantes

France pothesized that BARD1 is involved in the regulation ofcell death.

Summary Results

The BRCA1-associated protein BARD1 is a putative BARD1 Upregulation in Response to Hypoxia In Vivotumor suppressor. We suggest that BARD1 is a media- To investigate a putative role of BARD1 in cell death,tor of apoptosis since (1) cell death in vivo (ischemic we used a well-established mouse stroke model (Mar-stroke) and in vitro is accompanied by increased levels tinou et al., 1994). To induce ischemic stroke, medianof BARD1 protein and mRNA; (2) overexpression of cerebral artery occlusion (MCAO) was performed in theBARD1 induces cell death with all features of apopto- right brain hemisphere. Necrotic cell death occurs dur-sis; and (3) BARD1-repressed cells are defective for ing the first few days in the stroke region, while apopto-the apoptotic response to genotoxic stress. The pro- sis develops over several weeks in the surrounding tis-apoptotic activity of BARD1 involves binding to and sue, the penumbra (Martinou et al., 1994). In the brainelevations of p53. BRCA1 is not required for but par- of untreated animals, no BARD1 was found (Figure 1Aa),tially counteracts apoptosis induction by BARD1. A consistent with the previously described absence oftumor-associated mutation Q564H of BARD1 is defec- BARD1 mRNA in the central nervous system (Ayi et al.,tive in apoptosis induction, thus suggesting a role of 1998). Positive BARD1 immunostaining was detectedBARD1 in tumor suppression by mediating the signaling after ischemia predominantly in the penumbra (Figurefrom proapoptotic stress toward induction of apoptosis. 1Ac), increasing from day 1 up to day 14 in the ischemic

hemisphere (Figure 1B, left panel). In contrast, BRCA1Introduction was detected neither in the brain of untreated animals

nor of animals treated with MCAO (Figures 1Ab and 1B,BARD1 (BRCA1-associated RING domain 1) is described right panel). As positive control for the BRCA1 antibody,as a nuclear protein that associates with the breast cancer sections of mouse ovarian tissue were analyzed andsusceptibility gene product BRCA1 (Wu et al., 1996; Miki showed expression of BARD1 and BRCA1 in ovarianet al., 1994). The two proteins interact in vivo and in vitro follicles (Figure 1Ad–f). These results suggest that thethrough their respective RING finger domains (Wu et al., role of BARD1 in hypoxia-induced brain damage is inde-1996; Meza et al., 1999; Scully et al., 1997). As mutations pendent of BRCA1. BARD1 induction in ischemia wouldin the BRCA1 RING finger domain (Bienstock et al., 1996) be compatible with its involvement in cellular stress re-lead to increased cancer susceptibility, BARD1 was sponse and possibly the induction of apoptosis. To fur-thought to be implicated in BRCA1-mediated tumor sup- ther clarify the potential role of BARD1 in apoptosis, wepression. Consistent with this hypothesis, BARD1 muta- proceeded with in vitro studies.tions were found in breast, ovarian, and uterine tumors(Thai et al., 1998), and reduced expression of BARD1 BARD1 Expression in Responseis associated with spontaneous breast cancer cases to Genotoxic Stress In Vitro(Yoshikawa et al., 2000). The in vitro repression of BARD1 We first studied the response of various cell lines to pro-in murine mammary gland epithelial cells led to a premalig- apoptotic stress. Addition of the DNA-damaging chemo-

therapeutic agent doxorubicin to mouse mammary epithe-lial TAC-2 cells (Soriano et al., 1995) not only led to an7 Correspondence: [email protected]

Page 58: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Molecular Cell1256

Figure 1. BARD1 Expression Is Upregulated upon Stress

(A) BARD1 but not BRCA1 is upregulated in a mouse stroke model. In brains from untreated animals, neither BARD1 (a) nor BRCA1 (notshown) was detected by immunocytochemistry. As shown at day 7 after MCAO, BARD1 was not detected in the infarct core region (*) but inthe penumbra (c); staining remained negative for BRCA1 (b). Similar results were obtained at days 3 and 14. Bar, 550 �m. Unlike in the brain,BRCA1 (d) and BARD1 (e) are expressed in the mouse ovary. Primary antibody was omitted in (f). Bar, 100 �m.(B) As assessed by Western blotting, an increase of BARD1 can be observed from day 1 through day 14 after MCAO, in extracts from thetreated hemisphere (�) but not in the untreated hemisphere (�) (left panel). Expression of BARD1 and BRCA1 was compared on liver andnontreated (nt) or ischemic (isch) brain tissue (right panel). For both immunocytochemistry and Western blotting, the BARD1 antibody N-19(Santa Cruz sc-7373) was used. The same staining was observed with a C-terminal anti-BARD1 antibody (Santa Cruz sc-7372). Anti-BRCA1antibody was M-16 (Santa Cruz). Antibody preabsorption with specific peptides (Santa Cruz) or omission of primary antibody resulted inabsence of the signal (not shown).(C–E) BARD1 protein and mRNA is upregulated in cells exposed to genotoxic stress. Increase of p53 and BARD1 protein in TAC-2 cells,treated with doxorubicin (0, 5, or 10 �g/ml) for 12 hr, is demonstrated by Western blotting (C). Similarly, BARD1 expression is increased afterexposure to UV light for 1 to 4 min (2.5 to 10 Jm�2) (D). The expression of BARD1 was tested with anti-Bard1 antibody (PVC, Irminger-Fingeret al., 1998), and p53 expression was probed with anti-p53 antibody (FL-393, Santa Cruz). (E) RT-PCR of RNA from mouse ES cells exposedto UV light for 0, 1, or 4 min (0, 2.5, or 10 Jm�2) or treated with doxorubicin (10 �g/ml) shows an increase of BARD1 expression, while expressionof the control housekeeping gene �-tubulin remains constant.

Page 59: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

BARD1 Signaling toward p53-Mediated Apoptosis1257

Figure 2. BARD1 Overexpression Induces Apoptosis In Vitro

(A) TAC-2 cells were transfected with pcDNA3 vector sequences, BARD1, or the p53 sequences cloned into pcDNA3. Apoptosis inductionwas monitored by TUNEL staining 24 hr after transfection. Ten photographic fields were randomly chosen to evaluate the mean percentageof stained nuclei (bar chart).(B) DNA laddering was observed in cells transfected with BARD1 or p53 containing plasmids but not in cells transfected with pcDNA3. “Marker”indicates molecular weight marker smart ladder (Stratagene).(C) Caspase-3 activity assays were performed, and the fluorometric results of five independent transfection experiments are presented.Caspase-3 activity was significantly increased in BARD1 and p53-transfected cells.(D) BARD1 protein levels after doxorubicin treatment and after BARD1 transfection were compared in different cell lines on Western blots.Western blots were stripped and reprobed with anti-�-tubulin antibodies.(E) Comparison of mean expression levels of BARD1 in TAC-2 cells by flow cytometry after anti-BARD1 antibody staining, presented as xvalues of anti-BARD1-FITC. The increase of mean x values (arrow heads) in cells treated by doxorubicin or transfected by BARD1 as comparedto untreated cells is a direct measure of the increase in cellular BARD1 content. This increases of BARD1-derived fluorescence are similar indoxorubin-treated and BARD1-transfected cells.

Page 60: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Molecular Cell1258

Figure 3. Mapping of Proapoptotic Activity

BARD1 wild-type or mutant protein was expressed by transient transfection in mouse ES cells.Apoptosis induction in control transfectants and transfected cells was monitored 24 hr after transfection by DNA laddering (A) or TUNELassays, analyzed by flow cytometry (B). The expression after transfection was tested on Western blots of nontransfected (nt), BARD1-transfected, or BARD1Q564H (BARD1Q�H) mutant-transfected cells (C). An increase of 97 kDa BARD1 protein can be observed, whileexpression of actin is constant. Transfection efficiency was tested by cotransfection of GFP and was at least 20%.

increase in cellular p53 levels but also to marked elevations trol transfectants (empty vector or GFP) showed eitherno or only marginal signs of apoptosis. Thus, BARD1of cellular BARD1 levels (Figure 1C). Similar increase of

BARD1 in response to doxorubicin treatment was also transfection induced cell death, which displayed fea-tures of apoptosis. We also observed apoptosis inducedobserved in HeLa cells and human SH-SY5Y neuro-

blastoma cells (data not shown), as well as in mouse by BARD1 transfection in mouse ES, human SH-SY5Y,HeLa, and MCF-7 breast cancer cells, albeit to differentembryonic stem (ES) cells (see below). To test whether

cellular damage induced by means other than doxorubi- extents (data not shown). Thus, BARD1 is a potent in-ducer of apoptosis in cells of different tissue or speciescin also resulted in an increase of BARD1, cells of differ-

ent tissue origins were exposed to UV light. BARD1 was origins. The relative increase of BARD1 protein levelsobserved after transient transfection were comparableinduced in HeLa and TAC-2 cells (data not shown) and

mouse ES cells (Figure 1D) after UV treatment and fur- to the increase after doxorubicin treatment, as assessedby Western blotting (Figure 2D). However, since not allther incubation for 3 hr. The increase of BARD1 might

be regulated by transcriptional activation, as is demon- cells express BARD1 in a transient transfection assay,we also investigated BARD1 levels in individual cells bystrated by RT-PCR. Most strikingly in mouse ES cells,

no BARD1 mRNA could be detected in untreated cells, flow cytometry. Mean values of BARD1 levels in doxoru-bicin-treated cells were increased on average by a factorwhile BARD1 mRNA was clearly present after UV or

doxorubicin treatment (Figure 1E). These results sug- of 3.7 (� 2.4) with 5 �g/ml doxorubicin (Figure 2E), whilemean values for BARD1 levels in BARD1-transfectedgest that BARD1 upregulation upon genotoxic stress is

transcriptionally regulated. cells increased 2.5-fold (� 1.8) over controls, showing abroader distribution (Figure 2E). Thus, the mean cellularBARD1 levels are raised to comparable levels in doxoru-Elevated Expression of BARD1 Leads to Apoptosisbicin-treated and BARD1-transfected cells.To address the question of whether BARD1 upregulation

To study the mechanisms of proapoptotic signalingcauses cell death, we transfected TAC-2 cells with aby BARD1, we analyzed the induction of apoptosis byBARD1 expression plasmid. Attempts to establish sta-a BARD1 deletion mutant and the Q564H missense mu-ble BARD1-transfected cell lines were unsuccessful.tation (Figures 3A and 3B). TAC-2 cells or mouse ESOnly mock-transfected (empty vector or GFP) but notcells were transfected with pcDNA, BARD1, or theBARD1-transfected cells yielded viable cell lines, sug-BARD1Q564H mutant. Transfection efficiency was stan-gesting deleterious effects of BARD1 overexpression.dardized, and the protein expression levels of exoge-We therefore analyzed transiently transfected cells.nously expressed BARD1 and BARD1Q564H were com-Transient transfection of TAC-2 cells with BARD1 led toparable (Figure 3C). DNA laddering and TUNEL assaysapoptosis as assessed by TUNEL staining, DNA lad-were performed to monitor apoptosis induction. Inter-dering, caspase-3 activity tests (Figures 2A–2C), as wellestingly, the BARD1Q564H point mutation was largelyas annexin V staining and nuclear condensation moni-devoid of apoptosis-inducing activity. Since the Q564Htored by DAPI staining (data not shown). Transfectionmutation was shown to be associated with breast andof cells with the bona fide proapoptotic protein p53endometrial cancer (Thai et al., 1998), its reduced apo-(Sionov and Haupt, 1999) led to a similar extent of apo-

ptosis as transfection of BARD1 (Figures 2A–2C). Con- ptotic capacity provides a possible link between the

Page 61: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

BARD1 Signaling toward p53-Mediated Apoptosis1259

Figure 4. BARD1-Induced Apoptosis Is Independent of BRCA1

Transfection of pcDNA, BARD1, or BRCA1 expression plasmids was performed, and apoptosis induction was monitored 24 hr after transfection.Efficiency of transfection was 20% as determined by transfection of GFP expressing plasmids. DNA laddering was observed in BARD1-transfected cells but not in BRCA1- or pcDNA-transfected cells (A). Inhibition of BARD1-induced apoptosis by BRCA1 is presented incotransfection experiments in ES cells (B). Transient transfection assays were performed with pcDNA, BARD1, BARD1 � BRCA1, or BRCA1.The extent of apoptosis induction was measured by TUNEL staining and monitored by flow cytometry. Results presented are mean x (TUNEL-fluorescence) � range (n � 2).Expression of BARD1 and BRCA1 was monitored on Western blots (C). Gels were scanned, and increase of BARD1 and BRCA1 was measuredand were similar for both proteins.Influence of BRCA1 on apoptosis induction by BARD1 was analyzed in Brca1�/� and wild-type (wt) ES cells. DNA laddering was performedto qualitatively demonstrate apoptosis induction after doxorubicin treatment (5 �g/ml) or transient BARD1 transfection (D). Apoptosis inductionunder the same conditions was quantified by TUNEL assays of wild-type and Brca1�/� ES cells (E). Results presented are mean x (TUNEL-fluorescence) � range (n � 2).

tumor suppressor function of BARD1 and its proapo- 4B). Similar increase of expression was monitored forptotic activity. BRCA1 and BARD1 after transfection (Figure 4C). Inter-

estingly, the cotransfection of BRCA1 diminished ratherthan enhanced apoptosis induction by BARD1, as deter-Apoptosis Induction by BARD1 Is Independentmined by TUNEL staining and flow cytometry. To defini-of Functional BRCA1tively clarify whether apoptosis induction by BARD1 oc-An involvement of BRCA1 in apoptotic signaling hadcurs independently of BRCA1, we studied apoptosisbeen suggested previously (Harkin et al., 1999; Thangar-induction in Brca1-deficient mouse ES cells (Aprelikovaaju et al., 2000). Thus, the BARD1 effects might be medi-et al., 2001). As assessed by DNA laddering (Figure 4D),ated through or require BRCA1. To investigate this ques-BARD1 transfection induced apoptosis in Brca1-defi-tion further, we transfected BRCA1 into ES cells (Figurecient cells as efficiently as doxorubicin. These results4A) and into TAC-2 cells (data not shown); no apoptosisdemonstrate unequivocally that BRCA1 is not requiredoccurred under these conditions, while DNA ladderingfor BARD1-induced apoptosis.and positive TUNEL staining was observed after trans-

When apoptosis was assessed quantitatively byfection of BARD1 (Figure 4A). Thus, overexpression ofTUNEL assay, the Brca1-deficient cells clearly showedBRCA1 by itself does not induce apoptosis. To testincreased sensitivity to apoptosis induction by bothwhether BRCA1 synergized with BARD1 in apoptosis

induction, we cotransfected BRCA1 and BARD1 (Figure BARD1 and doxorubicin (Figure 4E). Thus, cotransfection

Page 62: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Molecular Cell1260

Figure 5. Repression of BARD1 Reduces the Apoptotic Response to Genotoxic Stress

BARD1 repression was achieved by stable transfection of BARD1 ribozyme in ES cells (ES riboBARD1) or a BARD1 antisense in murinemammary cells (TAC-2/ABI). After treatment with doxorubicin (5 �g/ml, 6 hr), apoptosis was monitored by flow cytometric analysis of TUNELstaining, as shown for ES cells and ES/riboBARD1 cells (A). Comparison of results obtained for ES, ES/riboBARD1, TAC-2, TAC-2/ABI, or ES/riboNox4 (transfected with an unrelated ribozyme) is shown in (B) with mean x (TUNEL-fluorescence) � SEM (n � 3).Cell numbers of TAC-2 cells or BARD1-repressed TAC-2/ABI cells in the absence or presence of doxorubicin (10 �g/ml) were monitored after2 and 4 days (C). Five fields of cells from each time point and each cell type were analyzed, and the mean � SEM remaining adherent cellsare shown (right panel). Caspase-3 activity (Molecular Probes) was measured by fluorometry after treatment with doxorubicin (10 �g/ml) forthe indicated periods of time (D). Results are mean � SEM (n � 3). Experiments were performed in the presence and absence of the caspase-3inhibitors DEVD.

Page 63: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

BARD1 Signaling toward p53-Mediated Apoptosis1261

Figure 6. BARD1-Induced Apoptosis Is Associated with p53 Accumulation

The p53 protein is upregulated in cells overexpressing BARD1 (A). Expression of p53 protein in cells harvested 24 hr after transfection with indicatedamounts of BARD1 and pcDNA3 or BARD1 was analyzed on Western blots (top panel). Effects of BARD1 and BARD1Q564H mutant (BARD1Q�H)on p53 expression were compared (middle and lower panels). While BARD1 increase is observed in BARD1- and BARDQ564H mutant-transfectedcells, significant increase of p53 is observed only in BARD1-transfected cells. The bar graph shows a comparison of p53 expression of cellstransfected with pcDNA3 or BARD1Q564H mutant (BARD1Q�H) as a percentage of p53 expression in BARD1-transfected cells.Increase of p53 in response to doxorubicin is compared in ES and TAC-2 cells, in BARD1-repressed ES/riboBARD1 and TAC-2/ABI cells, andin control transfectants ES/riboNox4 (B). Cells were incubated for 12 hr with doxorubicin (0, 5, or 10 �g/ml), and p53 and BARD1 accumulationwas monitored by Western blotting. Unchanged levels of actin are shown.p53 upregulation is not due to transcriptional upregulation (C). p53 mRNA expression is tested by RT-PCR of RNA prepared from untransfected,pcDNA-, or BARD1-transfected cells (left panel). Protein (�-p53) and mRNA (RTPCR) expression are compared in TAC-2 cells after the indicatedtime of incubation with doxorubicin (right panel). Control of independent gene expression was performed with primers against mouse �-tubulin.

of BRCA1 diminishes BARD1-induced apoptosis, and the exert its proapoptotic activity either as a homodimer orin a complex with other proteins.Brca1�/� genetic background enhances BARD1-induced

apoptosis. This suggests that BRCA1 actually antagonizesBARD1. Indeed, a BARD1 deletion mutant (RING finger Repression of BARD1 Reduces Apoptotic

Response to Genotoxic Stressdeletion of amino acids 1 through 237) retains apoptosisinducing activity (data not shown). Taken together, these So far we have demonstrated that BARD1 levels in-

crease in response to apoptotic stimuli and elevation ofresults suggest that the BRCA1/BARD1 complex is notinvolved in apoptosis induction and that BARD1 must BARD1 levels is sufficient to induce apoptosis. To as-

Page 64: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Molecular Cell1262

Figure 7. BARD1 Induces Apoptosis through Interaction with Functional p53

BARD1-p53 interaction was assayed by coimmunprecipitation (A). Anti-p53 (agarose-coupled; Santa Cruz) or anti-BARD1 antibodies (N-19;Santa Cruz) were used for immunoprecipitations (IP) from nontreated (non-treat) and doxorubicin-treated (doxo) cells, and Western blots wereincubated with anti-BARD1 (PVC, Irminger-Finger et al., 1998) and anti-PCNA (Santa Cruz) or anti-p53, respectively. Supernatants (S) andpellet (P) fractions are presented.Efficiency of BARD1-p53 interaction is compared in transient transfections of BARD1 or BARD1Q564H (BARD1Q�H) mutant (B). Western

Page 65: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

BARD1 Signaling toward p53-Mediated Apoptosis1263

sess whether BARD1 is required for the induction of mRNA after doxorubicin treatment were compared.While p53 protein levels were undetectable in untreatedapoptosis, we studied the effect of BARD1 repression.

For this purpose, we used two different systems for cells and increased after a few hours of treatment, p53mRNA levels, as detectable by RT-PCR, were elevatedBARD1 repression, ribozymes and antisense RNA.

Mouse ES cells stably expressing anti-BARD1 ribo- in untreated cells and did not increase significantly afterexposure to doxorubicin (Figure 6C, right panel). It iszymes were generated, and TAC-2 cells stably express-

ing BARD1 antisense RNA were described previously therefore unlikely that BARD1 and doxorubicin-inducedp53 protein accumulation are due to transcriptional(TAC-2/ABI; Irminger-Finger et al., 1998). In both sys-

tems, doxorubicin induced much less apoptosis (i.e., upregulation. It rather appears that BARD1 exerts itsaction on p53 at a posttranslational level, consistentTUNEL-positive cells), as compared to control cells or

cells transfected with an irrelevant ribozyme (Figures 5A with the concept of posttranslational regulation of p53levels during apoptosis (Lakin and Jackson, 1999).and 5B). Resistance to doxorubicin-induced apoptosis

through BARD1 antisense expression could also bemonitored by simple cell count (Figure 5C) or by measur- BARD1-p53 Interaction Is Requireding caspase-3 activity (Figure 5D). In control cells, doxo- for BARD1-Induced Apoptosisrubicin led to almost 100% cell death within a few days, To determine how BARD1 induces p53 accumulation,whereas BARD1-repressed cells showed delayed and we investigated whether the two proteins interact physi-reduced response. Taken together, these results sug- cally by performing coimmunoprecipitation experimentsgest that the apoptotic response to doxorubicin is, at (Figure 7A). Cell lysates from TAC-2 cells (Figures 7A–least in part, mediated through BARD1. 7C), mouse ES cells, or HeLa cells (data not shown),

untreated or after treatment with doxorubicin, weretested. When using either anti-BARD1 antibodies orApoptosis Induction by BARD1 Affects

p53 Protein Accumulation anti-p53 antibodies, BARD1 and p53 coimmunoprecipi-tated. Using p53 antibodies, 36% (�4; n � 2) of BARD1To further study the mechanism of BARD1-induced apo-

ptosis, we investigated the relationship between BARD1 was found in the pellet after treatment with doxorubicin(Figure 7A, upper panel), and p53 coimmunprecipitationand the p53 pathway of apoptosis. Transfection of

BARD1 into TAC-2 cells, which express wild-type p53 with BARD1 antibodies resulted in 52% (�3; n � 2)of the protein in the pellet fraction after doxorubicin(see Experimental Procedures; RT-PCR), led to a

marked increase in p53 protein levels (Figure 6A), signifi- treatment (Figure 7A, lower panel). The nuclear proteinPCNA, shown previously to be associated with BRCA1cantly higher than after transfection of vector sequences

(Figure 6A) or expression of unrelated proteins (data not in DNA repair functions (Scully et al., 1997) did not coim-munoprecipitate under our experimental conditions,shown). Importantly, transfection of the BARD1 mutation

Q564H, which showed reduced apoptotic capacity when demonstrating specificity of the p53-BARD1 interaction.However, this coimmunoprecipitation was observedexpressed in TAC-2 cells or ES cells (Figure 3), resulted

in significantly less p53 accumulation than that obtained only in doxorubicin-treated cells. The lack of coimmuno-precipiation of BARD1 and p53 in nontreated cells maywith wild-type BARD1 (Figure 6A, middle panel).

If BARD1 were a relevant link between DNA damage simply be due to the very low p53 protein levels in non-treated cells. The interaction between BARD1 and p53and p53 elevations, a decreased expression of p53 in

BARD1-suppressed cells could be expected. Indeed, was also observed in TAC-2 cells in which p53 elevationswere induced by BARD1 transfection (Figure 7B). How-BARD1 repression led to reduced p53 elevations in dox-

orubicin-treated cells, as shown for the two cell lines ever, coimmunoprecipitation was markedly decreasedfor the BARD1 mutant Q564H (Figure 7B). Only 19% (�2;described above (Figure 6B, top and middle panels).

Transfection of an irrelevant ribozyme (Nox4) did not n � 3) of p53 coimmunprecipitated with BARD1Q564H,while cells transfected with wild-type BARD1 showedalter p53 induction (Figure 6B, lower panel). The upregu-

lation of p53 observed at the protein level could be 58% of p53 protein in the pellet fraction. Thus, the tumor-associated BARD1Q564H mutant interacts less avidlydue to transcriptional upregulation of p53. RT-PCR was

performed to determine the changes in the p53 mRNA with p53 than wild-type BARD1.To further ascertain the role of p53 in BARD1-inducedlevels. No significant differences, either of p53 mRNA

or of the control housekeeping gene �-tubulin, were apoptosis, we investigated the apoptotic activity of BARD1in p53-deficient mouse ES cells (Corbet et al., 1999). Trans-observed in mock-transfected or pcDNA-transfected

cells when compared to BARD1-transfected cells (Fig- fection of BARD1 into p53�/� cells did not lead to apopto-sis, as determined by annexin V staining (Figure 7C) andure 6C, left panel). To further analyze p53 accumulation,

the time course of expression of p53 protein and p53 DNA ladder assay (Figure 7D), while apoptosis was in-

blots were probed with anti-p53 and anti-BARD1 antibodies. BARD1-induced apoptosis depends on functional p53, as demonstrated byannexin V staining (C) and DNA ladder assay (D) of wild-type or p53�/� ES cells transiently transfected with pcDNA3 or BARD1. BARD1 isrequired for p53 stabilization through protein-protein interaction as demonstrated by cotransfection of BARD1 and p53 in p53-deficient EScells (E); p53�/� ES cells were transfected with p53 or BARD1 plus p53. Increase of p53 was monitored on Western blots, and p53 proteinbands were quantified. Bar graph presents intensities of band as arbitrary units.Model of dual functionality of BARD1 (F). Increase of BARD1 upon stress: pathway (1) BRCA1-BARD1 heterodimer formation, observed asnuclear dots (Scully et al., 1997), induction of DNA repair functions; pathway (2) BARD1 excess over BRCA1 (possibly as homodimers), BARD1interaction with p53, affecting p53 stability and induction of apoptosis.

Page 66: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Molecular Cell1264

duced in the corresponding wild-type cells. Thus, p53 in Brca1-deficient embryos contribute to embryonic le-thality? Likewise, there is a high frequency of p53 muta-is required for BARD1-induced apoptosis.

The data presented so far demonstrate that increase tions in tumors with BRCA1 mutations (Buller et al., 2001;Greenblatt et al., 2001; Reedy et al., 2001). Thus, isof BARD1 leads to an increase of p53 protein, which is

not accompanied by detectable elevations of p53 the p53-dependent proapoptotic activity of BARD1 anobstacle toward malignant progression in cells withmRNA. Together with the direct protein-protein interac-

tion of the two molecules shown above, this suggests BRCA1 mutations, which has to be overcome throughp53 mutations?posttranslational stabilization of p53 by BARD1. To fur-

ther substantiate this point, we transfected p53-defi-cient ES cells with vector sequence, p53 plus control Mechanisms of BARD1-Induced p53 Elevationsvector, or p53 plus BARD1. In this system, p53 (ex- Our results demonstrate that DNA damage leads topressed from the constitutive CMV promoter) cannot be upregulation of BARD1, presumably by altering RNAcontrolled on a transcriptional level, but p53 protein levels (Figure 1E). Increase of BARD1 protein subse-stabilization must result from interaction with BARD1. quently affects p53 protein stabilization through a mech-Indeed, as shown in Figure 7E, markedly higher levels anism that most likely involves direct BARD1-p53 pro-of p53 protein are observed in cells cotransfected with tein-protein interaction. What could be the precisep53 and BARD1 than in cells transfected with p53 only. biochemical mechanisms of the BARD1-induced p53

upregulation? The most straightforward explanationwould be BARD1-induced changes of the p53 moleculeDiscussionleading to protein stabilization. Alternatively, however, itmight be due to an interference of BARD1 with ubiquitin-Our results demonstrate a function of BARD1 in signaling

between genotoxic stress and induction of apoptosis: dependent p53 degradation. Indeed, p53 turnover iscontrolled by the cellular protein degradation machineryBARD1 levels increase in response to genotoxic stress

and are necessary and sufficient for p53 upregulation that involving ubiquitin ligases, such as E6-AP (Scheffner etal., 1993; Rodriguez et al., 2000) and MDM-2 (Honda etresults in apoptotic response (Sionov and Haupt, 1999).

There are several interesting parallels between our re- al., 1997). Several observations point toward a role ofBARD1 in ubiquitination: BARD1 possesses an N-termi-sults and previously published observations. First, we

show decreased sensitivity to apoptosis-inducing drugs nal RING finger domain found in many proteins impli-cated in ubiquitination (Freemont, 2000), and as recentlyin BARD1-repressed cells (Figures 5A–5D), reminiscent

of the aberrantly reduced expression levels of BARD1 shown, the BARD1/BRCA1 complex possesses ubiqui-tin ligase activity (Scully and Livingston, 2000; Hashi-reported in spontaneous breast tumors (Yoshikawa et

al., 2000). Second, we show elevations of BARD1 during zume et al., 2001). Our data do not exclude that BARD1alters p53 ubiquitination; however, as BARD1 inductionapoptosis, reminiscent of BARD1 found in apoptotic

tumor cells (Gautier et al., 2000). Third, we show that of apoptosis is BRCA1 independent, ubiquitinationwould have to occur through mechanisms other thanthe BARD1Q564H mutant has a decreased proapoptotic

activity (Figure 3B), matching with its association with those described (Hashizume et al., 2001). Also, it is veryunlikely that the recently described function of BARD1breast and endometrium cancer (Thai et al., 1998).in inhibition of mRNA 3 end processing (Kleiman andManley, 2001) is linked to the BARD1-induced apoptosisRelationship between BRCA1 and BARD1-pathway described here, since it involves the formationInduced Apoptosisof a BARD1/BRCA1/CstF-50 complex.Hitherto described BARD1 functions depend on its as-

sociation with BRCA1, and a function of BRCA1 in apo-ptosis induction was proposed previously (Harkin et al., Dual Mode Hypothesis of BARD1 Function

Based on the BRCA1-dependent BARD1 functions de-1999). However, the function of BARD1 as apoptosisinducer, described here, has to be attributed to BARD1 scribed previously and on the BRCA1-indepedent

BARD1 function described in this paper, we propose aindependently of its association with BRCA1, because(1) only BARD1 but not BRCA1 increased after stroke; “dual mode of action” model of BARD1 function (Figure

7F). In the survival mode, BARD1 acts in a heterodimeric(2) the cotransfection of BRCA1 did not enhanceBARD1-induced apoptosis; and (3) BARD1 induced apo- complex with BRCA1 and is involved in DNA repair and

cell survival. In the death mode, BARD1 acts indepen-ptosis in BRCA1-deficient cells. Thus, BRCA1 is notrequired for apoptosis induction by BARD1. Our results dently of BRCA1 and induces apoptosis. Thus, ac-

cording to this model, the ratio of BRCA1 and BARD1even indicate that the interaction of BRCA1 with BARD1reduces apoptosis induction, as BARD1-induced apo- expressed in a given cell would determine cell fate; a

high BRCA1/BARD1 ratio would lead to DNA repair andptosis is diminished by cotransfection of BRCA1 butenhanced in BRCA1-deficient cells. survival, while a low ratio would lead to cell death. This

model would be compatible with the observed expres-The BRCA1-dependent decrease of BARD1-inducedapoptosis raises some intriguing questions concerning sion of BARD1 but not BRCA1 under conditions where

cell death occurs, e.g., stroke (see above) and the estrusrecently published results. Brca1-deficient mice are em-bryonic lethals (reviewed in Hakem et al., 1998) but can phase of the uterine cycle (Irminger-Finger et al., 1998).

It would also be compatible with the observed develop-be rescued through elimination of p53 (Ludwig et al.,1997). Similarly, Brca1 delta11/delta11 mice are rescued ment of a premalignant phenotype in vitro (Irminger-

Finger et al., 1998) and tumorigenesis (Yoshikawa et al.,from lethality by elimination of one p53 allele (Xu et al.,2001). Thus, does the proapoptotic activity of BARD1 2000) under conditions of reduced BARD1 levels.

Page 67: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

BARD1 Signaling toward p53-Mediated Apoptosis1265

Experimental Procedures primers were designed to amplify region 1840 through 2360 of theBARD1 cDNA, and p53 primers were designed to amplify region118 through 310 of the p53 cDNA sequence. The p53 coding se-Cerebral Ischemia

Cerebral ischemia was induced in the right brain hemisphere of quence was cloned by RT-PCR from TAC-2 cells (GenBank acces-sion number AY044188).C57BL/6 mice by medium cerebral artery occlusion (MCAO) as de-

scribed (Martinou et al., 1994). Mice were sacrificed at 1, 3, 7, and14 days after MCAO, and their brains were prepared for immunohis- Quantification of Apoptosistochemistry and Western blots. Protein extracts from brain tissue Apoptosis was monitored by several criteria, including changes inof the right and left hemisphere were prepared separately. cellular morphology observed by light microscopy and nuclear mor-

phology observed after DAPI staining; DNA laddering was performedwith Quick Apoptotic DNA Laddering Kit (Biosource International),Western Blotting and Immunoprecipitationannexin V staining and TUNEL assays were performed with kits fromFor Western blots, protein extracts from tissues or cell cultures wereMolecular Probes, and analyses were carried out on DAKO Galaxyprepared in standard lysis buffer (150 mM NaCl, 0.1% SDS, 50 mMpro flow cytometer. Caspase-3 activity tests were performed withTris [pH 8.0], and 2 mM EDTA [pH 8.0]) and analyzed by Westernkits from Molecular Probes and analyzed on a multilabel counterblotting using anti-p53 (Santa Cruz FL-393), anti-BARD1 PVC (Ir-Victor from Perkin Elmer (Switzerland).minger-Finger et al., 1998), N-19, and C-20 (Santa-Cruz sc-7373),

or anti-�-tubulin (Beohringer) antibodies. Immunoprecipitation wasAcknowledgmentsperformed by homogenizing cell samples in 100 �l (100�150 �g

total protein) RIPA buffer (150 mM NaCl, 1% NP-40, 0.5% NaDOC,This work was supported by a Johnson and Johnson Focused Giving0.1% SDS, 50 mM Tris [pH 8.0], and 2 mM EDTA [pH 8.0]). Cellgrant to I.I.F., grants from the AETAS Foundation to I.I.F. and K.H.K.,extracts were incubated with agarose-coupled p53 antibodies (Fland grants from the Swiss National Science Foundation to K.H.K.,393 Santa Cruz) in a 1 to 100 dilution for 2 hr, or BARD1 N-19M.D.D. (NCCR), and R.M. We are grateful to J.-C. Irminger, M. Mer-antibodies in 1 to 100 dilution overnight, followed by incubation withcola, and B. Vercher for critical comments on the manuscript. Wesepharose-bound protein G for 2 hr. Extracts were centrifuged, andare indebted to P. Vallet for MCAO surgery on mice, to F. Gautierpellets were washed twice with RIPA buffer and resuspended infor generating the BARD1 point mutation Q564H, and to N. Molnarfi100 �l RIPA buffer. 25 �l of supernatant and pellet fractions werefor technical support. In particular, we thank Beverly Koller for pro-analyzed by Western blotting. Total protein content in supernatantviding Brca1�/�and David Lane for providing p53�/� embryonic stemand pellet fractions was compared by Ponceau S staining aftercells.protein transfer and was 90% to 10% at least. Antibody reactive

bands were visualized, scanned, and quantified using Imagequant.Received March 13, 2001; revised July 11, 2001.

Doxorubicin Treatment and UV ExposureReferencesTAC-2, ES, or HeLa cells were grown to subconfluence in 100 mm

Petri dishes, exposed to UV light to reach 2.5 to 10 Jm�2, andAprelikova, O., Pace, A.J., Fang, B., Koller, B.H., and Liu, E.T. (2001).cultured for another 3 hr at 37C before harvesting. Doxorubicin (10,BRCA1 is a selective co-activator of 14–3-3 sigma gene expression50, 100 �g/ml) was added to the culture medium for 6 to 12 hrin mouse embryonic stem cells. J. Biol. Chem. 276, 25647–25650.before harvesting for apoptosis tests.Ayi, T.C., Tsan, J.T., Hwang, L.Y., Bowcock, A.M., and Baer, R.

Flow Cytometry (1998). Conservation of function and primary structure in the BRCA1-Cells were permeabilized and fixed with CytoFix (Pharmingen), and associated RING domain (BARD1) protein. Oncogene 17, 2143–incubated with anti-BARD1 antibodies and FITC-coupled secondary 2148.antibodies for flow cytometric analysis on a Galaxy Pro flow cyto- Bienstock, R.J., Darden, T., Wiseman, R., Pedersen, L., and Barrett,meter (Dako, Switzerland). J.C. (1996). Molecular modeling of the amino-terminal zinc ring do-

main of BRCA1. Cancer Res. 56, 2539–2545.Generation of Expression Clones and Cell Lines Buller, R.E., Lallas, T.A., Shahin, M.S., Sood, A.K., Hatterman-Zogg,BARD1 expression clone was produced by cloning the coding se- M., Anderson, B., Sorosky, J.I., and Kirby, P.A. (2001). The p53quence of the mouse BARD1 cDNA (Irminger-Finger et al., 1998) mutational spectrum associated with BRCA1 mutant ovarian cancer.into the KpnI and NotI sites of pcDNA3. The Q564H mutation was Clin. Cancer Res. 7, 831–838.generated with Quick Change Site-directed mutagenesis kit (Stra-

Corbet, S.W., Clarke, A.R., Gledhill, S., and Wyllie, A.H. (1999). P53-tagene). The murine BRCA1 cDNA, a generous gift of G. Lenoir, wasdependent and -independent links between DNA-damage, apopto-cloned into the pcDNA3 vector.sis and mutation frequency in ES cells. Oncogene 18, 1537–1544.Transfections were performed reproducibly with transfecten (Qia-Ferenczy, A. (1993). Ultrastructure of the normal menstrual cycle.gen). 1 to 4 �g of DNA was transfected per 100 mm Petri dish.Microsc. Res. Tech. 25, 91–105.Transfection efficiency was between 20 and 30% as assayed by

parallel transfection of GFP cloned into pcDNA3 and analyzed by Freemont, P.S. (2000). RING for destruction? Curr. Biol. 10, 84–87.flow cytometry and fluorescence microscopy. Gautier, F., Irminger-Finger, I., Gregoire, M., Meflah, K., and Harb,

J. (2000). Identification of an apoptotic cleavage product of BARD1Generation of BARD1 Antisense and Ribozyme Cell Lines as an autoantigen: a potential factor in the antitumoral responseThe ribozyme-carrying mouse ES cells were generated by polyclonal mediated by apoptotic bodies. Cancer Res. 60, 6895–6900.selection for neomycin resistance, using a BARD1-specific ribozyme Greenblatt, M.S., Chappuis, P.O., Bond, J.P., Hamel, N., and(Irminger-Finger et al., 1998). As control for nonspecific ribozyme Foulkes, W.D. (2001). TP53 mutations in breast cancer associatedeffects, mouse ES cells were transfected with an irrelevant ribo- with BRCA1 or BRCA2 germ-line mutations: distinctive spectrumzyme (riboNox4). The mouse mammary gland cell clone TAC-2/ABI, and structural distribution. Cancer Res. 61, 4092–4097.which expresses BARD1 antisense RNA (Irminger-Finger et al., 1998)

Hakem, R., de la Pompa, J.L., and Mak, T.W. (1998). Developmentalalso reduces BARD1 expression by 50 to 60%. The p53-deficient ESstudies of Brca1 and Brca2 knock-out mice. J. Mammary Glandcells were a generous gift of Alain Clarke and were described beforeBiol. Neoplasia. 7, 431–445.(Lakin and Jackson, 1999). Brca1-deficient ES cells were generouslyHarkin, D.P., Bean, J.M., Miklos, D., Song, Y.H., Truong, V.B., Englert,provided by the laboratory of B. Koller (Aprelikova et al., 2001).C., Christians, F.C., Ellisen, L.W., Maheswaran, S., Oliner, J.D., andHaber, D.A. (1999). Induction of GADD45 and JNK/SAPK-dependentRT-PCRapoptosis following inducible expression of BRCA1. Cell 97,Total RNA was prepared from cell cultures with Qiagen purification575–586.kits. RT-PCR (Perkin Elmer) was performed in simultaneous one-

step or in two-step reactions, leading to identical results. BARD1 Hashizume, R., Fukuda, M., Maeda, I., Nishikawa, H., Oyake, D.,

Page 68: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Molecular Cell1266

Yabuki, Y., Ogata, H., and Ohta, T. (2001). The RING heterodimer Identification of a RING protein that can interact in vivo with theBRCA1 gene product. Nat. Genet. 14, 430–440.BRCA1-BARD1 is an ubiquitin ligase inactivated by a breast cancer-

derived mutation. J. Biol. Chem. 276, 14537–14540. Xu, X., Qiao, W., Linke, S.P., Cao, L., Li, W.M., Furth, P.A., Harris,C.C., and Deng, C.X. (2001). Genetic interactions between tumorHonda, R., Tanaka, H., and Yasuda, H. (1997). Oncoprotein MDM2suppressors Brca1 and p53 in apoptosis, cell cycle and tumorigene-is a ubiquitin ligase E3 for tumor suppressor p53. FEBS Lett. 420,sis. Nat. Genet. 28, 266–271.25–27.

Yoshikawa, K., Ogawa, T., Baer, R., Hemmi, H., Honda, K., Yamauchi,Irminger-Finger, I., Soriano, J.V., Vaudan, G., Montesano, R., andA., Inamoto, T., Ko, K., Yazumi, S., Motoda, H., et al. (2000). AbnormalSappino, A.P. (1998). In vitro repression of Brca1-associated RINGexpression of BRCA1 and BRCA1-interactive DNA-repair proteinsdomain gene, Bard1, induces phenotypic changes in mammary epi-in breast carcinomas. Int. J. Cancer 88, 28–36.thelial cells. J. Cell Biol. 143, 1329–1339.

Kleiman, F.E., and Manley, J.L. (1999). Functional interaction ofAccession NumbersBRCA1-associated BARD1 with polyadenylation factor CstF-50.

Science 285, 1576–1579.The accession number for p53 cDNA expressed in mouse mammary

Kleiman, F.E., and Manley, J.L. (2001). The BARD1-CstF-50 interac- epithelial cell line TAC-2 is AY044188.tion links mRNA 3 end formation to DNA damage and tumor sup-pression. Cell 104, 743–753.

Lakin, N.D., and Jackson, S.P. (1999). Regulation of p53 in responseto DNA damage. Oncogene. 18, 7644–7655.

Ludwig, T., Chapman, D.L., Papaioannou, V.E., and Efstratiadis, A.(1997). Targeted mutations of breast cancer susceptibility gene ho-mologs in mice: lethal phenotypes ofBrca1, Brca2, Brca1/Brca2,Brca1/p53, and Brca2/p53 nullizygous embryos. Genes Dev. 11,1226–1241.

Martinou, J.C., Dubois-Dauphin, M., Staple, J.K., Rodriguez, I.,Frankowski, H., Missotten, M., Albertini, P., Talabot, D., Catsicas,S., Pietra, C., et al. (1994). Overexpression of BCL-2 in transgenicmice protects neurons from naturally occurring cell death and exper-imental ischemia. Neuron 13, 1017–1030.

Meza, J.E., Brzovic, P.S., King, M.C., and Klevit, R.E. (1999). Mappingthe functional domains of BRCA1. Interaction of the ring finger do-mains of BRCA1 and BARD1. J. Biol. Chem. 274, 5659–5665.

Miki, Y., Swensen, J., Shattuck-Eidens, D., Futreal, P.A., Harshman,K., Tavtigian, S., Liu, Q., Cochran, C., Bennett, L.M., Ding, W., etal. (1994). A strong candidate for the breast and ovarian cancersusceptibility gene BRCA1. Science 266, 66–71.

Reedy, M.B., Hang, T., Gallion, H., Arnold, S., and Smith, S.A. (2001).Antisense inhibition of BRCA1 expression and molecular analysisof hereditary tumors indicate that functional inactivation of the p53DNA damage response pathway is required for BRCA-associatedtumorigenesis. Gynecol. Oncol. 81, 441–446.

Rodriguez, M.S., Desterro, J.M., Lain, S., Lane, D.P., and Hay, R.T.(2000). Multiple C-terminal lysine residues target p53 for ubiquitin-proteasome-mediated degradation. Mol. Cell. Biol. 20, 8458–8467.

Scheffner, M., Huibregtse, J.M., Vierstra, R.D., and Howley, P.M.(1993). The HPV-16 E6 and E6-AP complex functions as a ubiquitin-protein ligase in the ubiquitination of p53. Cell 75, 495–505.

Scully, R., and Livingston, D.M. (2000). In search of the tumour-suppressor functions of BRCA1 and BRCA2. Nature 408, 429–432.

Scully, R., Chen, J., Ochs, R.L., Keegan, K., Hoekstra, M., Feunteun,J., and Livingston, D.M. (1997). Dynamic changes of BRCA1 sub-nuclear location and phosphorylation state are initiated by DNAdamage. Cell 90, 425–435.

Sionov, R.V., and Haupt, Y. (1999). The cellular response to p53: thedecision between life and death. Oncogene 18, 6145–6157.

Soriano, J.V., Pepper, M.S., Nakamura, T., Orci, L., and Montesano,R. (1995). Hepatocyte growth factor stimulates extensive develop-ment of branching duct-like structures by cloned mammary glandepithelial cells. J. Cell Sci. 108, 413–430.

Thai, T.H., Du, F., Tsan, J.T., Jin, Y., Phung, A., Spillman, M.A.,Massa, H.F., Muller, C.Y., Ashfaq, R., Mathis, J.M., et al. (1998).Mutations in the BRCA1-associated RING domain (BARD1) gene inprimary breast, ovarian and uterine cancers. Hum. Mol. Genet. 7,195–202.

Thangaraju, M., Kaufmann, S.H., and Couch, F.J. (2000). BRCA1facilitates stress-induced apoptosis in breast and ovarian cancercell lines. J. Biol. Chem. 275, 33487–33496.

Wu, L.C., Wang, Z.W., Tsan, J.T., Spillman, M.A., Phung, A., Xu,X.L., Yang, M.C., Hwang, L.Y., Bowcock, A.M., and Baer, R. (1996).

Page 69: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

60

2. Nuclear-cytoplasmic translocation of BARD1 is linked to its apoptotic activity Jefford CE, Feki A, Harb J, Krause KH, Irminger-Finger I. Oncogene. 2004 Apr 29;

23(20):3509-20.

The work performed in this report is a continuation of the previous paper in that it

defines the region necessary for induction of apoptosis. According to our previous

results, we know that BARD1 can induce apoptosis in transient transfection

experiments. In order to identify the BARD1 protein domain or region responsible for

triggering the apoptotic response, we constructed a series of deletion mutants of

BARD1. We then analysed their capacity to induce apoptosis by TUNEL assay, DNA

laddering assays and morphometric measurements and assessed their nuclear

localization by immunohistochemistry, by epi-fluorescence microscopy of GFP-

tagged BARD1 deletion recombinants. We demonstrated in particular that the

minimal region required for inducing apoptosis maps to a region that does not

contain a functional domain, which lies in between the ankyrin repeats and the

BRCT domain. Surprisingly, this region, comprising the amino acids 510-604,

coincides with 2 presumed cancer predisposition mutations C557S and Q564H and

correlates with the tumour suppressor function of BARD1. In contrast with previous

reports which stipulated an exclusive localization of BARD1, we showed that BARD1

localizes to the cytoplasm in vitro and in vivo. Furthermore, we showed by using a

BARD1-EGFP fusion and BARD1-EGFP deletion proteins that BARD1 localizes to

the cytoplasm in spite of several predicted nuclear localizing signal sequences

(NLSs). Finally, we showed that apoptosis is concomitant with enhanced

cytoplasmic localization of BARD1, as well as the appearance of a p67 cleavage

product. The electron microscopy (EM) images were obtained by Jean Harb, and the

western blot and the immunohistochemistry performed by Anis Feki.

Page 70: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Nuclear–cytoplasmic translocation of BARD1 is linked to its apoptotic

activity

Charles Edward Jefford1, Anis Feki1, Jean Harb2, Karl-Heinz Krause1 and IrmgardIrminger-Finger*,1

1Biology of Aging Laboratory, Department of Geriatrics, University of Geneva, 1225 Geneva, Switzerland; 2Institut de Biologie,INSERM U419, 9 quai Moncousu, 44035 Nantes Cedex 03, France

The tumor suppressor protein BARD1 plays a dual role inresponse to genotoxic stress: DNA repair as a BARD1–BRCA1 heterodimer and induction of apoptosis in aBRCA1-independent manner. We have constructed aseries of BARD1 deletion mutants and analysed theircellular distribution and capacity to induce apoptosis. Asopposed to previous studies suggesting an exclusivelynuclear localization of BARD1, we found, both in tissuesand cell cultures, nuclear and cytoplasmic localization ofBARD1. Enhanced cytoplasmic localization of BARD1,as well as appearance of a 67 kDa C-terminal proteolyticcleavage product, coincided with apoptosis. BARD1translocates to the nucleus independently of BRCA1.For recruitment to nuclear dots, however, the BRCA1-interacting RING finger domain is required but notsufficient. Protein levels of N-terminal RING fingerdeletion mutants were much higher than those of full-length BARD1, despite comparable mRNA levels, sug-gesting that the N-terminal region comprising the RINGfinger is important for BARD1 degradation. Sequencesrequired for apoptosis induction were mapped between theankyrin repeats and the BRCT domains coinciding withtwo known cancer-associated missense mutations. Wesuggest that nuclear and cytoplasmic localization ofBARD1 reflect its dual function and that the increasedcytoplasmic localization of BARD1 is associated withapoptosis.Oncogene (2004) 23, 3509–3520. doi:10.1038/sj.onc.1207427Published online 12 April 2004

Keywords: BARD1; BRCA1; apoptosis; nuclear export;protein stability; RING

Introduction

A function of BARD1 in a pathway of stress-inducedsignaling towards apoptosis was discovered previously(Irminger-Finger et al., 2001). Overexpression ofBARD1 inducing apoptosis in a number of cell lines,and genotoxic stress triggering apoptosis is accompanied

by upregulation of BARD1. Thus, BARD1 serves as amediator between proapoptotic stress signals and p53-dependent apoptosis. In line with its tumor suppressorfunction, depletion of BARD1 protein in vitro throughantisense BARD1 mRNA or ribozyme expression,annihilates p53 upregulation and apoptotic response tostress and leads to genomic instability and a premalig-nant phenotype (Irminger-Finger et al., 1998, 2001;Soriano et al., 2000). BARD1 is implicated in a numberof other functions most of them dependent on itsinteraction with the tumor suppressor protein BRCA1,of which it is the major binding partner in a yeast two-hybrid screen (Wu et al., 1996).

BRCA1 and BARD1 heterodimerize through theirrespective N-terminal RING finger domains (Meza et al.,1999). Association of the C64G mutation within the N-terminal RING finger of BRCA1 with familial breastand ovarian cancer (Wu et al., 1996; Ruffner et al., 2001)suggests a relevant function for the BRCA1–BARD1interaction in tumor suppression. Structural analysis ofthe BARD1–BRCA1 heterodimeric complex by NMR(Brzovic et al., 2001b) and mutational analysis demon-strated that abolishing heterodimer formation (Morriset al., 2002) results in genomic instability, uncontrolledcell proliferation and tumorigenesis (Baer, 2001; Parvin,2001).

Evidence for BARD1/BRCA1 heterodimer formationin vivo comes from the observation that BARD1colocalizes with BRCA1 and the repair protein Rad51in nuclear dots during S phase (Jin et al., 1997) and tonuclear foci in response to DNA damage (Scully et al.,1997b). Mutations in the RING finger of BRCA1,disrupting BRCA1/BARD1 interactions, annihilate theformation of nuclear foci (Chiba and Parvin, 2002;Fabbro et al., 2002), which suggests that the interactionwith BARD1 is required for this localization.

A recent report shows that BARD1 plays a role aschaperone for the correct translocation of BRCA1 intothe nucleus (Fabbro et al., 2002). Deletion of theBRCA1-RING domain leads to exclusion of BRCA1from the nucleus, and a BRCA1 mutant lacking NLS,but retaining the RING domain structure, is translo-cated to the nucleus by BARD1 overexpression. Mostimportantly, the binding of BARD1 inhibits the nuclearexport of BRCA1 by masking its nuclear export signal(NES) (Brzovic et al., 2001a; Fabbro et al., 2002). Sincemore than 75% of all BRCA1 is complexed with

Received 22 July 2003; revised 20 November 2003; accepted 2 December2003; Published online 12 April 2004

*Correspondence: I Irminger-Finger;E-mail: [email protected]

Oncogene (2004) 23, 3509–3520& 2004 Nature Publishing Group All rights reserved 0950-9232/04 $25.00

www.nature.com/onc

Page 71: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

BARD1 (Baer and Ludwig, 2002), BARD1 plays animportant role in trapping BRCA1 within the nucleus.Whether BARD1 itself depends on other proteins forentering and exiting the nucleus has not beendetermined.

The BRCA1/BARD1 heterodimer exhibits ubiquitinligase activity, which is abolished by mutations withinthe RING finger domain of BRCA1 (Hashizume et al.,2001; Chen et al., 2002). BARD1 supports the associa-tion of BRCA1 with Polymerase II holoenzyme, asuspected target of BARD1/BRCA1 ubiquitylation,which is abrogated by a deletion of the BRCA1-RINGfinger domain (Chiba and Parvin, 2002). Furthermore,the stability of BARD1 and BRCA1 proteins themselvesmight be controlled by ubiquitylation (Mallery et al.,2002; Xia et al., 2003). Besides degradation, ubiquityla-tion can act as a regulator of nuclear–cytoplasmicshuttling, as described for the nuclear export of p53 (Guet al., 2001; Lohrum et al., 2001). A similar mechanismmight exist in the case of BARD1 and BRCA1, inparticular since BARD1 and BRCA1 become ubiquiti-nylated upon association (Chen et al., 2002). It wassuggested that nuclear–cytoplasmic shuttling has func-tional implications since BARD1 and BRCA1 are lessprone for degradation when bound together but theirdissociation after ubiquitylation could lead to nuclearexport and degradation (Fabbro and Henderson, 2003).

In addition to its interaction with BRCA1, BARD1associates with a number of other proteins involved indifferent functions (Irminger-Finger and Leung, 2002),all of which imply a nuclear localization such as mRNApolyadenylation factor CstF50 (Kleiman and Manley,1999; Kleiman and Manley, 2001), the NFK-B regulatoroncoprotein Bcl-3 (Dechend et al., 1999), the complexcontaining BRCA1 and CtIP (Yu and Baer, 2000), andthe Ewing’s sarcoma protein EWS (Spahn et al., 2002).

BARD1 structurally relates to BRCA1, as it containsN-terminal RING finger and C-terminal BRCT do-mains (Wu et al., 1996; Ayi et al., 1998); however, novelfunctions of BARD1 might relate to its ankyrin repeats,which are found in proteins with very diverse functions,and to regions without known protein motifs. Here, wereport that BARD1 is found in the nucleus and thecytoplasm and that its apoptotic function is associatedwith its cytoplasmic localization. Furthermore, bymapping the regions responsible for intracellular loca-lization or required for its apoptotic function, using anenhanced green fluorescent protein (EGFP), tagged tothe C-terminal end of full-length BARD1 or variousdeletion variants, we demonstrate their respectivedynamic intracellular localization and the correlationof cytoplasmic localization and apoptotic activity.

Results

Cytoplasmic localization of BARD1

The tissues with the highest expression of BARD1 andBRCA1 mRNAs are spleen and testis (Wu et al., 1996;Ayi et al., 1998; Irminger-Finger et al., 1998). In the

spleen, a BRCA1-associated function of BARD1 couldbe expected, as well as a function in apoptosis. Weapplied immunohistochemistry to detect the BARD1protein in the spleen. BARD1 staining was observed inthe red pulp and the white pulp, both regions known forthe high turnover of lymphocytes (Figure 1a). Theintracellular localization of BARD1 was both nuclearand cytoplasmic and in some cells exclusivelycytoplasmic (Figure 1a).

It was interesting to determine the intracellularlocalization of BARD1 in tissues where BARD1 isupregulated in response to cellular stress. In the brain,no BARD1 expression can be found normally butmaximum BARD1 expression is observed in the regionsadjacent to the stroke induced by ischemia (Irminger-Finger et al., 2001), in the penumbra, where apoptosishas been described (Ferrer and Planas, 2003). Immuno-histochemisty on mouse brain sections after ischemiademonstrates that the intracellular localization ofBARD1 is rather cytoplasmic than nuclear (Figure 1b).

Figure 1 In vivo localization of BARD1. (a) BARD1 staining onsections of mouse spleen using anti BARD1 N-19 and PVCantibodies. BARD1 can be localized to the red pulp and white pulp(magnification � 10). Intracellular localization is observed withboth antibodies. Bar 100mm. (b) BARD1 staining (anti-BARD1 N-19) is shown in a mouse brain after ischemia. Penumbra region isshown in overview in (� 4.5). BARD1 staining is cytoplasmic inmost cases, as exemplified by arrows in higher magnifications in(� 40). Bar 100mm

Nuclear–cytoplasmic translocation of BARD1CE Jefford et al

3510

Oncogene

Page 72: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Strong evidence for cytoplasmic localization of BARD1protein is the observed staining of protrusions ofneuronal cells (Figure 1b). Control staining withoutprimary antibody or in the presence of antigenic peptidedid not produce staining (not shown). Antibodiesagainst the N-terminal and the C-terminal portion ofthe protein present the same distribution. These datasuggest that in vivo, apoptosis, occurring physiologicallyor induced upon stress, is accompanied by a cytoplasmiclocalization of BARD1, and that the apoptotic functionof BARD1 might be confined to the cytoplasm.

Subcellular distribution of BARD1 in vitro

Previous reports stipulate that BARD1 is found innuclear extracts and localizes to BRCA1 foci in thenucleus during the S phase (Jin et al., 1997; Scully et al.,1997a) and to distinct intranuclear foci upon genotoxicstress (Scully et al., 1997a). To explore the dynamics ofBARD1 localization at various stages of the cell cycleand apoptosis, immunofluorescent staining and electronmicroscopy (EM) were performed on NIH 3T3 cells(Figure 2). Immunoreactivity with anti-BARD1 anti-bodies results in a punctuate staining in the nucleus ofinterphase cells but diffuse staining in the cytoplasm aswell (Figure 2a). Similar results were obtained whenBARD1 protein localization was assayed in other celllines (TAC2, HEK 293T, HeLa) and when otherantibodies (N-19, C-20, H300) were used (Figure 2b).Enhanced staining of BARD1 in nuclear dots isobtained by confocal microscopy. By EM, immunor-eactive dots of BARD1 were identified in the nucleusand with similar frequency in the cytoplasm (Figure 2c)corroborating the data obtained by immunofluores-cence. Interestingly, we also observed that cells progres-sing through mitosis accumulate higher levels ofBARD1 than G1 or S-phase cells. BARD1 is excludedfrom the condensed chromatin during metaphase–anaphase stages of mitosis (Figure 2a). This localizationof BARD1 contrasts with the reported binding ofBRCA1 to the centrosomes during mitosis (Hsu andWhite, 1998; Deng, 2002; Lotti et al., 2002). Our datasuggest that BARD1 might have a specific function inthe cytoplasm, and that BARD1 is not engaged infunctions associated with the centrosomal location ofBRCA1.

Intracellular localization of BARD1-EGFP and BARD1-EGFP deletion mutants

Mouse BARD1 has three putative NLS sequenceslocated close to or within each functional domain(Figure 3a): NLS1 -RKKNSIKMWFSPRSKKV- start-ing at position 130, NLS2 -PARKRNH- at 408 andNLS3 -RKPK- at 693. When applying Reinhardt’smethod (http://psort.nibb.ac.jp/) for the prediction ofnuclear localization with a reliability coefficient of94.1%, the prediction for intracellular localization formouse BARD1 has a 52.2% score for the nucleus,30.4% for the mitochondria, 13% for the cytoplasm anda 4.3% for the cytoskeleton. Together the cytosolic

fraction values add up to 47.8% implying a possible rolefor BARD1 within the cytoplasm. Human BARD1 hassix putative NLS sequences and a higher predictionscore for the nucleus of 82.6%.

In order to ascertain BARD1’s intracellular localiza-tion and to determine which domains of BARD1 orwhich NLS are necessary for nuclear import, wegenerated either EGFP-tagged full-length mouseBARD1 or deletion mutants (Figure 3a). Constitutiveexpression of EGFP-BARD1 fusion proteins wasobtained under the transcriptional control of the CMVpromoter. To determine the localization of BARD1-EGFP or deletion mutants, we transfected the BARD1-EGFP plasmids in various cell lines and monitored theirlocalization by fluorescent microscopy at different timepoints (12–30 h) after transfection. The intracellular

Figure 2 Cytoplasmic and nuclear localization of BARD1 in vitro.(a) Indirect immunofluorescent staining of BARD1 protein inNIH3T3 cells using anti-BARD antibody (PVC). Interphase cellsshow immunoreactive dots in the nucleus and the diffuse staining inthe cytoplasm. In mitotic cells, BARD1 is excluded from thechromatin. Nuclear staining with DAPI of identical field is shown.Bar 10 mm. (b) Immunofluorescent staining of HeLa cells withvarious anti-BARD1 antibodies (PVC, C-20, N-19, H-300) shownuclear dots and cytoplasmic staining. (c) EM photomicrograph ofan interphase cell. BARD1 reactive dots are evenly dispersed in thecytoplasm and the nucleus. EM image of an early mitotic cell showsthat BARD1 reactive dots are excluded from the condensedchromatin

Nuclear–cytoplasmic translocation of BARD1CE Jefford et al

3511

Oncogene

Page 73: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

distribution of all BARD1-EGFP fusion proteinsdiffered greatly from the nuclear and cytoplasmiclocalization of EGFP alone (Figure 3b). The analysisof a number of 8–12 microscopic fields from at leastthree independent transfection experiments at 24 h aftertransfection (Figure 3c), showed that full-lengthBARD1-EGFP locates to the nucleus more frequently(76%) than to the cytoplasm (24% of the cases).Surprisingly, the D-RING-EGFP and D-HIND-EGFP,which lack the first NLS at position 130, but also lackthe BRCA1-interacting RING finger domain, exhibiteda most pronounced preference for the nucleus. Interest-ingly, D-RING-EGFP, which localized exclusively tothe nucleus, never aggregates within nuclear dots duringthe S phase. The RING domain by itself when fused toEGFP localizes evenly to the cytoplasm and nucleus, asdid ANK-EGFP (Figure 3b). Both constructs have onlya single NLS. Most interestingly, the C-terminaldeletion, D-BRCT-EGFP, localized almost exclusivelyto the nucleus and formed distinct speckles within thenucleus, as did the full-length BARD1-EGFP. Thesedata indicate that the nuclear localization of BARD1 isindependent of BRCA1, since the BARD1 deletionconstructs lacking the RING finger have even increasednuclear localization. However, the aggregation withinnuclear dots might depend on BRCA1, since neither D-RING-EGFP nor D-HIND-EGFP, or ANK-EGFPconcentrate on nuclear dots. The N-terminal regioncontaining the BRCA1 interaction domain might berequired but not sufficient for BARD1 localization tonuclear dots.

Mapping of apoptotic function of BARD1

While the function of the conserved RING fingerdomain is well described, less contributions have beenmade to elucidate the functional significance of theankyrin repeats, the BRCT domains, or the interveningsequences. To determine the proapoptotic region ofBARD1 and to investigate whether any of the conservedsequence motifs was involved, we analysed the inductionof apoptosis by BARD1 deletion mutants missing eitherRING finger, ankyrin, or BRCT domain (Figure 4).Maximal induction of apoptosis, as tested by DNAladdering (Figure 4a) and TUNEL staining (Figure 4b),was observed only with the full-length BARD1 con-struct. Efficient induction of apoptosis was observedwith a variety of deletion mutants (Figure 4c), namelydeletion of the BARD1 C-terminal BRCT domains (D-BRCT), of the region comprising the ankyrin repeats (D-ANK), of the N-terminal domain comprising the RINGfinger (D-RING), or the deletion of N-terminal and theC-terminal regions (ANK2). One mutant was devoid ofapoptosis-inducing activity, namely of the constructexpressing the BARD1 N-terminus, RING. From theseexperiments, it can be concluded that the regionsufficient for apoptosis induction includes sequencesfrom amino acid 316 through 604 and that a minimalregion required for apoptosis induction comprisessequences from amino acid 510 through 604.

Figure 3 Subcellular localization of EGFP-tagged BARD1deletion mutants. (a) Schematic representation of mouse BARD1and BARD1-EGFP fusion proteins. Relative positions of mouseBARD1 functional domains (RING, green, ANKYRIN, blue, andBRCT, red), and three NLS (blue) are indicated. Schematicdiagram of the various deletion mutants of BARD1 fused to theEGFP. (b) Merge of DNA stain DAPI and epifluorescent images ofHEK-293T shows subcellular localization of transfected BARD1-deletion mutants tagged with EGFP 30h post transfection. (c)Intracellular localization of BARD1 and deletion mutants. 293-Tcells were transfected with BARD1-EGFP mutants and fixed inPFA 2%. Histograms show distribution of cells displaying eithernuclear (N), nuclearþ cytoplasmic (NC), or cytoplasmic localiza-tion. Statistics of localization of EGFP fluorescence was performedon eight different randomized fields under fluorescent microscope.Results from at least three different experiments were pooled togive significant numbers of transfected cells

Nuclear–cytoplasmic translocation of BARD1CE Jefford et al

3512

Oncogene

Page 74: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

To demonstrate that apoptosis induction was directlydue to the expression of BARD1, we determined theapoptosis inducing capacity of BARD1-EGFP andBARD1 deletion mutants fused to EGFP in transienttransfection assays in TAC-2 cells. With comparabletransfection efficiencies, the overexpression of theexogenous BARD1-EGFP proteins induces morpholo-gic changes and nuclear fragmentation, which is notobserved with cells transfected with control plasmidEGFP only or in nontransfected cells. All constructscomprising the minimal apoptotic region led to apop-tosis induction after transient transfection. Staining withDNA dye DAPI demonstrates that apoptotic cells,recognized by their nuclear fragmentation, are clearlycorrelated with cells expressing EGFP fusion proteins inthe case of BARD1-EGFP, D-RING-EGFP, D-BRCT-EGFP, and ANK-EGFP, while expression of EGFP, orRING-EGFP was not correlated with apoptosis(Figure 5). Inspection of random fields showed that at

early time points after transfection (24–48 h) thefrequency of apoptosis was lower for D-RING-EGFP,D-HIND-EGFP, and D-BRCT-EGFP than for otherdeletion mutants or for BARD1-EGFP, and at all timesinsignificantly low for EGFP and RING-EGFP.

Cytoplasmic localization of BARD1 during apoptosis

Since endogenous and exogenous expression of BARD1is found in both the nucleus and cytoplasm, it wasspeculated that the apoptotic function of BARD1 wascorrelated with its cytoplasmic localization as observedfor other gene products of the apoptotic pathway(Conus et al., 2000). When individual cells weremonitored after transient transfection with BARD1-EGFP, it was observed that the EGFP-tagged BARD1,or truncated BARD1, D-RING-EGFP, was translo-cated from the nucleus to the cytoplasm duringapoptosis (Figure 6a). This was specifically pronounced

Figure 4 Mapping of the apoptotic domain of BARD1. Apoptosis induction, as observed 48 h post transfection in TAC2 cells, wasmonitored of BARD1 and deletion mutants. (a) BARD1 and BARD1-deletion mutants induce DNA laddering and apoptosis, exceptdeletion mutant RING lacking aa 316–777 (excluding the ANKYRIN and BRCT domains). (b) Apoptotic cells were quantified byTUNEL assay. FITC-positive cells were counted by flow cytometry. (c) Schematic presentation of BARD1 and regions expressed ondeletion constructs is presented. Red bracket indicates minimal required region for apoptosis induction. Position of cancer associatedmissense mutations are indicated (Thai et al., 1998; Ghimenti et al., 2002)

Nuclear–cytoplasmic translocation of BARD1CE Jefford et al

3513

Oncogene

Page 75: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

in cells transfected with D-RING-EGFP that aftertransfection is located exclusively to the nucleus. Todemonstrate that exogenous BARD1 D-RING-EGFPand endogenous BARD1 were both translocated fromthe nucleus to the cytoplasm during apoptosis, immuno-fluorescence microscopy was performed on cells trans-fected transiently with D-RING-EGFP. Staining withanti-BARD1 antibodies (PVC) directed against theRING domain (Irminger-Finger et al., 1998) demon-strates that D-RING-EGFP, not recognized by anti-BARD1-PVC, first localized to the nucleus, then at laterstages of apoptosis is translocated to the cytoplasm,while endogenous BARD1, detected with anti-BARD1antibody PVC, is still localized to nuclear dots. At latestages of apoptosis, both exogenous and endogenousBARD1 are found mostly excluded from the nucleus(Figure 6b). This observation is consistent with the

hypothesis that excess of BARD1 over BRCA1 couldlead to cytoplasmic localization of BARD1 and initiationof the apoptotic pathway (Irminger-Finger et al., 2001).

To further address the apoptosis-related translocationof BARD1, its distribution between nucleus andcytoplasm was assayed after treatment with the apop-tosis inducing drug doxorubicin, both by immunofluor-escence microscopy and Western blotting. MCF7 cellswere analysed with or without doxorubicin treatment. Inuntreated cells, BARD1 protein levels are very low buthigher in the nucleus than in the cytoplasm; uponapoptosis, overall BARD1 levels are increased, but thedistribution of BARD1 between the nucleus andcytoplasm is inversed (Figure 7a). Interestingly, BRCA1remains localized to nuclear dots after doxorubicintreatment, as observed by staining with anti-BRCA1antibodies (Figure 7a).

Figure 5 Apoptosis induction by EGFP-tagged BARD1 and BARD1-deletion mutants. (a) Epifluorescence images, merged withDNA stain DAPI, of TAC-2 cells transfected with the full-length or the various deletion-bearing BARD1-EGFP plasmids. Cells werefixed and stained with DAPI 48 h after transfection. Constructs used for transfection were as described in Figure 3. Nuclearfragmentation in EGFP positive cells is observed for all BARD1 EGFP fusion constructs except for RING-EGFP. Apoptosisprogression of EGFP-positive cells was correlated with intensity of EGFP stain. (b) Percentage of GFP-positive apoptotic cells intranscient transfection assays. Six microscopic fields of two to three independent transfection assays were analysed. GFP-positiveapoptotic cells were determined based on nuclear morphology. Percentages of apoptotic cells are presented

Nuclear–cytoplasmic translocation of BARD1CE Jefford et al

3514

Oncogene

Page 76: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

To investigate BARD1 translocation after doxorubi-cin treatment biochemically, protein extracts from cellstreated with doxorubicin were prepared and nuclear andcytoplasmic fractions were analysed on Western blots(Figure 7b). To follow the translocation of BARD1 tothe cytoplasm after BARD1-induced apoptosis, wemonitored BARD1 expression in cytoplasmic andnuclear fractions at different time points after treatment.Using antibodies against different regions of BARD1,we observed that full-length 97 kDa BARD1 concentra-tions were increased at early time points after treatmentwith doxorubicin, and decreased at later time points.Analysis of nuclear and cytoplasmic extracts demon-strates that the concentration of BARD1 in the nucleusis increased during 6 h of doxorubicin treatment andthen decreased. In the cytoplasm, however, BARD1 of97 kDa is slightly increased and a protein of 67 kDa ismassively increased after 5 h of treatment. These datacould be explained with the following hypothesis:BARD1 levels increase after apoptosis induction, butexcess BARD1 translocates to the cytoplasm and iscleaved and gives rise to the stable C-terminal 67 kDafragment. The appearance of a C-terminal 67 kDaBARD1 cleavage product is consistent with the previousreport that a truncated 67 kDa product of BARD1

accumulates during apoptosis due to cleavage by thecaspase-dependent kinase calpain (Gautier et al., 2000).

Apoptotic function of BARD1 regulated through proteinstability

It was observed that BARD1-EGFP and EGFP-taggeddeletion constructs did not give rise to identical proteinexpression levels of the fusion products after transfec-tion. To investigate whether the differences observedwere due to differences in transcriptional activity, theregulation of translation, or protein stability, wemonitored the expression of BARD1-EGFP and ofBARD1 deletion mutants on the level of mRNA andprotein expression (Figure 8a–c). RT–PCR was per-formed using primers directed against EGFP and

Figure 6 BARD1-EGFP translocates to the cytoplasm duringapoptosis. (a) NIH3T3 cells transfected with D-RING-EGFP aremonitored at different time points after transfection. FITCfluorescence is extruded from the nucleus. (b) Comparison ofendogenous and exogenous EGFP-tagged BARD1 location, D-RING-EGFP mutant is presented. Endogenous BARD1 wasdetected with anti-BARD1 antibody PVC reacting with theBARD1 N-terminal region (Irminger-Finger et al., 1998), whichis omitted in D-RING-EGFP. Merge of EGFP signal (FITC) andanti-BARD1 signal (rhodamin) images shows that both theexogenous and the endogenous BARD1 are translocated fromthe nucleus to the cytoplasm in cells that are undergoing apoptosis.An early (arrow) and a late (arrow head) apoptotic cell are shown

Figure 7 Endogenous BARD1 translocates to the cytoplasmduring apoptosis. (a) BARD1 staining is shown in MCF-7 cells.BARD1 is localized primarily to the nucleus in untreated cells andtranslocates to the cytoplasm after UV exposure. Note that lowlevels of BARD1 in untreated cells were visualized by overexposureof the image. UV exposure was as described previously (Irminger-Finger et al., 2001). BRCA1 staining was performed with anti-BRCA1 antibody 5-MO. (b) Western blot analysis of nuclear versuscytoplasmic extracts in NIH3T3 cells following exposure todoxorubicin in a time-dependent manner shows increased expres-sion of BARD1 during apoptosis and demonstrates the appearanceof a different truncated form of BARD1 found in the cytoplasm.Antibodies used were C-20 (Santa Cruz) recognizing a C-termepitope (** represents the 97 kDa isoform and * the 67 kDa.)

Nuclear–cytoplasmic translocation of BARD1CE Jefford et al

3515

Oncogene

Page 77: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

against BARD1. The mRNA expression levels of full-length BARD1-EGFP and of the diverse BARD1-EGFP deletion mutants were compared by usingprimers either covering different regions of the mouseBARD1 coding sequence or directed against the GFPportion of the fusion proteins. The mRNA levels weresimilar in all observed cases (Figure 8a), indicating thatmRNA expression levels were directly correlated to theconcentration of plasmid DNA transfected.

To test protein translation potential of the variousconstructs, in vitro transcription/translation was per-formed to monitor protein stability in vitro. Full-lengthBARD1 and deletion mutants were transcribed andtranslated into protein with similar efficiency, althoughslightly elevated levels of protein were obtained with theD-RING construct (Figure 8b). However, protein

expression levels after transfection of EGFP-taggedconstructs appeared to vary. Western blots were probedwith antibodies against GFP in cell extract collected 24 hafter transfection. Probing of protein extracts aftertransfection of the various BARD1 expression con-structs exhibited significant differences in expressionlevels. Highest expression was observed for D-RING-EGFP, less for D-HIND-EGFP and ANK-EGFP, verylittle for full-length BARD1-EGFP, which was compen-sated by increasing the protein concentration forBARD1-EGFP loaded on the gel (Figure 8c). Repeat-edly, a C-terminal cleavage product could be detectedmore readily with anti-GFP antibodies than full-lengthBARD1-EGFP, suggesting that BARD1 and BARD1deletion mutants that comprise the N-terminal region ofBARD1 are prone for degradation.

Discussion

Nuclear–cytoplasmic distribution of BARD1

The general idea that BARD1 is a nuclear protein isbased exclusively on in vitro studies. We have investi-gated the localization of BARD1 in tissues. In parti-cular, we were interested to determine the intracellularlocalization of BARD1 in tissues where apoptosisoccurs, either physiologically or induced. Our resultsdemonstrate that BARD1 is localized to both thenucleus and the cytoplasm in vivo. We further exploredthe localization of BARD1 in a variety of cell linesin vitro. Both immunofluorescence and electromicro-scopy revealed that endogenous BARD1 resides in thenucleus and the cytoplasm. Using EGFP-taggedBARD1 and BARD1 deletion mutants, we confirmedthat all BARD1 deletion proteins were translocated tothe nucleus, but a variable proportion of the proteinswas also detected in the cytoplasm. These data contrastwith previous reports showing that BARD1 is localizedexclusively to the nucleus colocalizing with BRCA1 indiscreet dots during the S phase (Scully et al., 1997a). Itis possible that in human cells with human BARD1,which has six NLS, the propensity of nuclear localiza-tion is higher than that for mouse BARD1, which hasonly three putative NLS. All BARD1-EGFP fusionproteins, described here, are translocated to the nucleus,indicating that a single NLS sequence is sufficient for thecorrect addressing of BARD1 to the nucleus. Thedifferential intracellular localization of the EGFP-tagged constructs may be correlated with the numberof NLS present within each specific construct. Interest-ingly, the BARD1-RING finger motif is not required fornuclear localization, and nuclear localization of BARD1is therefore independent of BRCA1. In fact, it has beenshown that on the contrary BRCA1 requires BARD1for entering and remaining in the nucleus, and this isdependent on the BRCA1-RING domain (Fabbro et al.,2002). Furthermore, we demonstrate that the BRCT-deletion mutant of BARD1 localizes to specific nucleardots, as does full-length BARD1, but not the RINGdomain by itself, suggesting that RING is not sufficientfor localization of BARD1 in nuclear dots.

Figure 8 Differential stability of BARD1 and BARD1 deletionproteins. (a) RT–PCR analysis of cell extracts after transfectionwith plasmids as indicated. Primers were against two differentregions of BARD1 or against EGFP. Similar amounts of mRNAwere amplified from cells transfected with the various transcripts.The RING-EGFP construct can only be detected with the EGFPprimers. To visualize loading differences, two different volumeswere loaded. (b) In vitro transcription and translation of BARD1and deletion bearing constructs. Plasmids were transcribed andtranslated (with S-35 labeled methionine) in vitro, and analysed byPAGE. All constructs gave rise to comparable amounts of protein.(c) Western blot analysis of whole-cell extracts of cells transfectedwith the aforementioned BARD1-EGFP-fusion plasmids andprobed with an anti-EGFP (TP-401) antibody. Expression of fulllength was reduced compared to N-terminal deletion mutants andwas loading was fivefold increased. Full-length BARD1-EGFP andD-RING-EGFP give rise to a degradation product that comigrateswith D-HIND-EGFP. Triangles indicate proteins with expectedsize for each fusion protein on the left (for EGFP and ANK-EGFP) and on the right side (for D-HIND-EGFP, BARD1-EGFP,and D-RING-EGFP)

Nuclear–cytoplasmic translocation of BARD1CE Jefford et al

3516

Oncogene

Page 78: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

An alternative localization of BARD1 occurs duringapoptosis. BARD1 and a truncated form of BARD1,p67 BARD1, appears in the cytoplasm during apopto-sis. This isoform of BARD1 represents the C-terminalportion of the protein, recognized by antibodies directedagainst 50 C-terminal amino acids. A 67 kDa C-terminaltruncated form of BARD1 was reported previously tobe generated during apoptosis in various cell linesthrough cleavage by the caspase-dependent proteasecalpain (Gautier et al., 2000). The p67 protein observedin the cytoplasmic fraction is consistent with theaccumulation of the p67 BARD1 cleavage product inapoptotic tumor cells (Gautier et al., 2000).

Apoptosis is triggered by overexpression of a minimalrequired region of BARD1

Overexpression of all forms of BARD1-EGFP anddeletion mutants, with the exception of the RING-EGFP construct, induced apoptosis, based on TUNELand DNA fragmentation assays, and visualized by thecondensed and fragmented nuclear morphology ascompared to the EGFP control. Furthermore, usingdeletion mutants of BARD1, a minimal region requiredfor triggering apoptosis could be determined. Asobserved with the overexpression of full-length BARD1,cells transfected with BARD1-EGFP, or BARD1-deletion mutants cannot be cultured over time due tothe apoptotic effect of the fusion proteins. The minimalregion required for the apoptotic function comprises thesequence between the end of the ankyrin repeats and thebeginning of the BRCT domain. Interestingly, thisregion harbors two point mutations (C557S andQ564H) associated with tumors of the breast, ovary,and uterus (Thai et al., 1998; Ghimenti et al., 2002). Inline with these findings, we have previously reportedthat the mutant Q564H when used in transienttransfections is less efficient in apoptosis induction thanfull-length BARD1 (Irminger-Finger et al., 2001). Takentogether, these data suggest that the primary tumorsuppressor function of BARD1 is linked to its BRCA1-independent function in inducing apoptosis.

Protein stability is increased in the absence of the RINGdomain

When expression of BARD1-EGFP or the tagged fusionproteins was monitored on Western blots, the proteinlevels varied according to the specific plasmid constructand despite identical transfection conditions and iden-tical plasmid DNA concentrations used. Interestingly,BARD1-EGFP, RING-EGFP and D-BRCT-EGFPconstructs showed increased susceptibility for degrada-tion. However, BARD1 protein stability appears to beincreased in the absence of the RING domain.Consistent with this observation, all BARD1-EGFPdeletion-mutants that bear the RING finger domainhave lower expression levels than the mutants lackingthe RING domain. One possible explanation for thedegradation of these fusion proteins is the reportedubiquitylation activity of the RING domain (Chen et al.,

2002). Possibly, the BARD1-RING domain is a targetfor protein degradation by the ubiquitin pathway, asspeculated for BRCA1 (Fabbro and Henderson, 2003),and BARD1–BRCA1 may target excess BARD1 fordegradation, but BARD1–BRCA1 dimer formation pro-tects from degradation. Furthermore, it was also hypo-thesized that ubiquitylation could act as nuclear exportsignal, as it is the case for p53 (Gu et al., 2001; Lohrumet al., 2001), and that BARD1 influences the nuclearexport of BRCA1, which is controlled by ubiquitylation(Fabbro and Henderson, 2003). The observation ofBARD1 translocation to the cytoplasm, together withits described degradation, suggests that BARD1 trans-location could be based on the ubiquitylation, specifi-cally since BARD1 does not have an NES on its own.

The model derived from these observations

It was suggested that the BRCA1/BARD1/PolymeraseII holoenzyme complex is the sensor for double-stranded DNA breaks while arresting the cell in S phasefor repair mechanisms to occur (Parvin, 2001). Accord-ing to our model (Figure 9), BARD1 may be the keymolecule that could trigger the cell-cycle block as aBARD1–BRCA1 complex and apoptosis through anincrease of BARD1. Initiation of apoptosis is accom-panied by BARD1 cleavage and accumulation of astable C-terminal cleavage product, which might repre-sent an irreversible decision for apoptosis. BRCA1 mayrecruit BARD1 to activate ubiquitylation of the Poly-merase II holoenzyme, and provoking its degradationthrough the proteasome pathway. BRCA1 was sug-gested to be liberated from BARD1 at later stages(Fabbro and Henderson, 2003), although it is not clearwhat triggers the separation of BARD1 and BRCA1. Incase of a high ratio of BARD1/BRCA1, the excess

Figure 9 Model of BARD1 cytoplasmic translocation. BARD1enables BRCA1 to be transported to the nucleus. Excess BARD1 istranslocating to the cytoplasm more readily. However, full-lengthBARD1 might be exported and degraded. BARD1 degradationdepends on the presence of the RING structure. A C-terminalfragment, p67 BARD1, presumably due to proteolytic cleavage, isstable and found in the cytoplasm. p67 BARD1 is no bindingpartner for BRCA1 but has apoptotic activity

Nuclear–cytoplasmic translocation of BARD1CE Jefford et al

3517

Oncogene

Page 79: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

BARD1 is found in the cytoplasm and might bedegraded, but the cleavage mechanism that eliminatesthe N-terminal sequences including the RING domainproduces a stable C-terminal portion of BARD1. Thisp67 cleavage product of BARD1 is mimicked by thedeletion mutant D-RING. Since D-RING retainsapoptotic activity, it is likely that the cleaved form ofBARD1 has an apoptotic function. Most importantly,both the BARD1 cleavage product p67 and D-RING aredevoid of the RING finger and seem more resistantagainst degradation. Therefore the cleavage of BARD1during apoptosis could provide a positive regulatoryfeedback loop upon apoptosis induction.

In conclusion, BARD1 has at least five differentfunctions in binding to BRCA1, all of which aredisrupted by mutations that effect the interaction ofBRCA1with BARD1: (1) tumor suppression and geneticinstability (Wu et al., 1996; Ruffner et al., 2001), (2)colocalization in nuclear dots upon genotoxic stress(Scully et al., 1997a, b), (3) nuclear translocation ofBRCA1 (Fabbro et al., 2002), (4) ubiquitin ligaseactivity of the BARD1/BRCA1 heterodimer (Hashi-zume et al., 2001; Chen et al., 2002), (5) recruiting ofPolymerase II holoenzyme (Chiba and Parvin, 2002).The localization of BARD1 in nuclear dots requiresRING domain, but also the additional domains thatmight promote binding of other proteins than BRCA1,since the BARD1-RING domain by itself does notlocalize to nuclear dots. It could be concluded, that full-length BARD1 might be required for BRCA1-depen-dent functions associated with nuclear localization, buta C-terminal portion of BARD1 is sufficient forapoptotic activity, associated with cytoplasmic localiza-tion (Figure 9).

Materials and methods

Cell cultures

TAC-2 cells is a clonal subpopulation derived from theNMuMG normal murine mammary gland epithelial cell line(CRL 1636; American Type Culture Collection, Rockville,MD, USA). They were cultured in collagen-coated tissueculture dishes in high-glucose DMEM supplemented with 10%fetal calf serum (FCS), penicillin (110mg/ml) and streptomycin(110mg/ml). The Human Embryonic Kidney HEK 293T cellline, as well as HeLa cell lines, were obtained from theAmerican Tissue Type Culture Collection. NIH-3T3 andMCF-7 cell lines were purchased from the NIH (Bethesda,MD, USA). All cell lines were cultured in 10 cm plastic tissueculture dishes (NUNC, Roskilde, Denmark) and incubated at371C in humidified air containing 5% CO2. Adherentsubcultures were detached from culture dishes using a solutionof Trypsin/EDTA in HBSS without Ca2þ or Mg2þ . Subcon-fluent cell cultures were passaged regularly at 1 : 10 dilutiontwice a week. All cell culture media, solutions, buffers andantibiotics were purchased from GIBCO (Invitrogen AG,Basel, Switzerland).

Plasmid constructions

Two sets of deletion mutants were generated with and withoutthe EGFP fusion tag. The full-length mouse cBARD1

(Irminger-Finger et al., 1998) was used to generate thefollowing constructs in pcDNA3 behind a CMV promoter:(1) full-length BARD1 (aa 1–765); (2) DRING (aa 137–765)using the EcoRI(396) restriction site; (3) DBRCT (aa 1–604)ScaI(1812); (4) DANK (aa 1–316/510–765) Stu/Hpa; (5)ANK2(aa 316–604) StuI/ScaI; (6) RING (aa 1–266) EcoRI. Theplasmid pEGFP-N1 (Enhanced Green Fluorescent Protein)was purchased from Clontech and was used to generate eitherthe full-length BARD1-EGFP or BARD1 deletion bearingmutants fused in frame with the EGFP tag. The followinginserts were ligated in frame with the EGFP: (1); (2); (3); (6)and (7) DHIND (aa 394–765); (8) ANK (aa 316–604).

Transfection and whole-cell extract preparation

Cells were seeded in 10 cm Petri dishes and transfected with2mg of DNA using the Effectene reagents by Qiagen. Cellswere harvested at 18, 24 and 48 h post-transfection. They werecentrifuged and washed once in PBS, resuspended in 0.3ml ofice-cold radio immunoprecipitation assay (RIPA) buffer(150mM NaCl, 1% Nonidet P-40, 0.5% NaDOC, 0.1%SDS, 50mM Tris, pH 8.0, 2mM EDTA, pH 8.0) andsupplemented with protease inhibitors (1mM PMSF, 3 mg/mlaprotinin, 10 mg/ml leupeptin)). Cells were incubated on ice for30 min. The extracts were centrifuged at 13 000 g in amicrocentrifuge for 15min at 41C in order to remove celldebris and clear the cell lysate.

Nuclear and cytoplasmic fractions

For separation of cytosolic fraction, cells were incubated inLysis Buffer (Triton X-100 (1%), HEPES pH 7.6. (50 mM),NaCl (150mM), NaF (100 mM), Na-pyrophosphate (50 mM),EDTA (4mM), Na3VO4 (10mM)) supplemented with proteaseinhibitors (as above) and rotated at 41C for 30 min. Super-natant was separated from nuclear fractions by centrifugationat 13 000 r.p.m. for 15 min. The pellet was washed in the WashBuffer (Lysis Buffer þ 25% glycerol) and centrifuged oncemore as above. The supernatant wash was added to cytosolicfraction. Nuclear fraction was lysed in the nuclear lysis buffer(NLB) (Wash Buffer with 330mM NaCl), sonicated, incubatedon ice for 30min and centrifuged as above to separate celldebris from nuclear fraction.

Western blots

In vitro transcription/translation products were obtained usinga kit from Promega. Equal concentrations of all in vitrotranscription/translation, whole-cell extracts as well as cyto-solic and nuclear fractions per gel were loaded onto SDS–PAAgels according to Laemmli’s protocol and were transferred topolyvinylidene fluoride (PVDF) membranes. All primaryantibodies were incubated at room temparature (RT) atindicated dilutions for 1 h in TBS/5% nonfat dry milk andwashed five times 3min in TBS/0.1% Tween 20. SecondaryHRP-conjugated antibodies were incubated at a 1 : 2000dilution (Santa Cruz) at RT for 1 h, and bands were revealedusing the Enhanced Chemiluminescence (ECL) system byBoehringer.

Antibodies

The anti-BARD1 (N-19, C-20, H-300), and anti-p53 (FL-393)antibodies were purchased from Santa Cruz. Anti-BRCA1 (5-MO) was a generous gift from Dr A Vincent and G Lenoir.Anti-GFP (TP-401) was purchased from Chemokine (TorreyPines Biolabs, Houston, TX, USA)). PVC is a previouslydescribed anti-BARD1 antibody (Irminger-Finger et al., 1998).

Nuclear–cytoplasmic translocation of BARD1CE Jefford et al

3518

Oncogene

Page 80: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Secondary conjugated antibodies to horseradish peroxidase(HRP) were purchased from Santa Cruz and were compatiblewith Cruz Molecular Weight Markers (Santa Cruz, CA, USA).

Assessment of apoptosis

Cells were harvested 48 h post transfection. Trypsinized cellswere centrifuged and washed once with PBS prior assay.Quantification of apoptosis was measured by performingTUNEL assay, using the APO-DIRECT kit (Pharmingen).At least 10 000 events were measured on a flow cytometer fromDAKO corporation and analysed using the software packageFlowMax from PARTEC. Assessment of DNA fragmentationwas performed with Suicide-Track DNA isolation kit (Onco-gene) and DNA ladders were stained with ethidium bromideon 1% agarose gels.

Fluorescence microscopy

Cells were seeded on glass coverslips in 6-well culture dishes orplated on an 8-well coverslide and transfected or not with theaforementioned plasmids. Cells were washed in HBSS, andeither fixed 2% paraformaldehyde (PFA) in Hank’s balancedsalt solution (HBSS) 15min at RT or in methanol for 6min at�201C and rinsed in acetone for 30 s. Coverslides weremounted and observed under a Nikon epi-fluorescencemicroscope and images captured with a 3.3 mega-pixel CCDcamera and processed with MetaMorph software (Visitron).

Tissue preparation for immunohistochemistry

Two adult male mice Balb/c were used in this part of the study.This study was conducted according to the rules written by theSwiss veterinary office (Bern). The animals were killed bycervical dislocation. After the animals were killed, the thoracicaorta was cannulated and the vasculature flushed with sodiumchloride solution (0.9 g/l). The fixation was performed byperfusion with 2% glutaraldehyde and 2% paraformaldehydein 0.05 M phosphate buffer, pH 7.4.

Immunohistochemistry

The immersion-fixed mouse brain and spleen were embeddedin paraffin. Consecutive orthogonal 5 mm sections weremounted on polylysine-coated slides. Sections were counter-stained with hematoxylin/eosin (HM). The first section of eachset of sections was stained HM and used to determine tubulestages. The remaining sections were further processed for

immunocytochemical staining. The endogenous peroxidasewas quenched by 15-min incubation in 2% hydrogen peroxidein PBS. Unmasking of the epitopes was carried out by boilingthe deparaffinized rehydrated sections for 10min (two times,5 min each) in 10 mM citrate buffer, pH 6.0, using a microwaveoven at 5000W power output. Primary antibodies were used ina 1 : 50 dilution.

Immunogold labeling and silver enhancement for EM

Cells were fixed with 3% freshly depolymerized paraformalde-hyde and 0.15%. glutaraldehyde in 0.1 M phosphate buffer(PB) pH 7.2 for 1 h at room temperature. They were thenpermeabilized with PB containing 0.1% Triton X-100. Theimmunostaining was then performed using ultrasmall goldconjugates and R-Gent SE-EM silver enhancement reagentusing the Aurion Kit (Biovalley, Marne-la-Vallee, France) asindicated by the manufacturer. Primary antibody used was C-20 diluted at 1/100. After immunostaining, the cells werepostfixed for 1 h at 41C with osmium tetraoxide buffered at pH7.2 with symmetric collidin. After two washes in the samebuffer, the preparations were dehydrated through a series ofgraded ethanol and propylene oxide, and the pellets embeddedin Epon 812 medium. Ordinary thin sections were then cutwith a Reichert ultramicrotome, stained with uranyl acetateand lead citrate, or left unstained. All sections were thenexamined and photographed under a JEM electronmicroscope.

Note added in proof

Coincidentally, our results are corroborated by similar findings(Rodriguez et al., 2003) published recently in this journal asour article was undergoing final revision.

Acknowledgements

This work was supported by Swiss National Science Founda-tion grant to IIF and KHK, and Johnson & Johnson andAETAS awards to IIF. We are grateful to Dr Gilbert LeNoirand Dr Anne Vincent for the generous gift of the BRCA1antibody, Dr Monica Sauer for discussing with us unpublishedobservations, Dr Wai-Choi Leung for critical comments on themanuscript, and Aurelie Caillon for excellent technicalsupport.

References

Ayi TC, Tsan JT, Hwang LY, Bowcock AM and Baer R.(1998). Oncogene, 17, 2143–2148.

Baer R and Ludwig T. (2002). Curr. Opin. Genet. Dev., 12,

86–91.Baer R. (2001). Nat. Struct. Biol., 8, 822–824.Brzovic PS, Meza JE, King MC and Klevit RE. (2001a).

J. Biol. Chem., 276, 41399–41406.Brzovic PS, Rajagopal P, Hoyt DW, King MC and Klevit RE.

(2001b). Nat. Struct. Biol., 8, 833–837.Chen A, Kleiman FE, Manley JL, Ouchi T and Pan ZQ.

(2002). J. Biol. Chem., 277, 22085–22092.Chiba N and Parvin JD. (2002). Cancer Res., 62, 4222–4228.Conus S, Rosse T and Borner C. (2000). Cell Death Differ., 7,

947–954.Dechend R, Hirano F, Lehmann K, Heissmeyer V, Ansieau S,

Wulczyn FG, Scheidereit C and Leutz A. (1999). Oncogene,18, 3316–3323.

Deng CX. (2002). Oncogene, 21, 6222–6227.Fabbro M and Henderson BR. (2003). Exp. Cell Res., 282,

59–69.Fabbro M, Rodriguez JA, Baer R and Henderson BR. (2002).

J. Biol. Chem., 277, 21315–21324.Ferrer I and Planas AM. (2003). J. Neuropathol. Exp. Neurol.,

62, 329–339.Gautier F, Irminger-Finger I, Gregoire M, Meflah K and Harb

J. (2000). Cancer Res., 60, 6895–6900.Ghimenti C, Sensi E, Presciuttini S, Brunetti IM, Conte P,

Bevilacqua G and Caligo MA. (2002). Genes ChromosomesCancer, 33, 235–242.

Gu J, Nie L, Wiederschain D and Yuan ZM. (2001). Mol. Cell.Biol., 21, 8533–8546.

Hashizume R, Fukuda M, Maeda I, Nishikawa H, Oyake D,Yabuki Y, Ogata H and Ohta T. (2001). J. Biol. Chem., 276,

14537–14540.

Nuclear–cytoplasmic translocation of BARD1CE Jefford et al

3519

Oncogene

Page 81: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Hsu LC and White RL. (1998). Proc. Natl. Acad. Sci. USA, 95,

12983–12988.Irminger-Finger I and Leung WC. (2002). Int. J. Biochem. Cell

Biol., 34, 582–587.Irminger-Finger I, Leung WC, Li J, Dubois-Dauphin M, Harb

J, Feki A, Jefford CE, Soriano JV, Jaconi M, Montesano Rand Krause KH. (2001). Mol. Cell, 8, 1255–1266.

Irminger-Finger I, Soriano JV, Vaudan G, Montesano R andSappino AP. (1998). J. Cell Biol., 143, 1329–1339.

Jin Y, Xu XL, Yang MC, Wei F, Ayi TC, Bowcock AM andBaer R. (1997). Proc. Natl. Acad. Sci. USA, 94,

12075–12080.Kleiman FE and Manley JL. (1999). Science, 285, 1576–1579.Kleiman FE and Manley JL. (2001). Cell, 104, 743–753.Lohrum MA, Woods DB, Ludwig RL, Balint E and Vousden

KH. (2001). Mol. Cell. Biol., 21, 8521–8532.Lotti LV, Ottini L, D’Amico C, Gradini R, Cama A, Belleudi

F, Frati L, Torrisi MR and Mariani-Costantini R. (2002).Genes Chromosomes Cancer, 35, 193–203.

Mallery DL, Vandenberg CJ and Hiom K. (2002). EMBO J.,21, 6755–6762.

Meza JE, Brzovic PS, King MC and Klevit RE. (1999). J. Biol.Chem., 274, 5659–5665.

Morris JR, Keep NH and Solomon E. (2002). J. Biol. Chem.,277, 9382–9386.

Parvin JD. (2001). Proc. Natl. Acad. Sci. USA, 98, 5952–5954.Rodriguez JA, Schuchner S, Au WW, Fabbro M and

Henderson BR. (2003). Oncogene.Ruffner H, Joazeiro CA, Hemmati D, Hunter T and Verma

IM. (2001). Proc. Natl. Acad. Sci. USA, 98, 5134–5139.Scully R, Chen J, Ochs RL, Keegan K, Hoekstra M, Feunteun

J and Livingston DM. (1997a). Cell, 90, 425–435.Scully R, Chen J, Plug A, Xiao Y, Weaver D, Feunteun J,

Ashley T and Livingston DM. (1997b). Cell, 88, 265–275.Soriano JV, Irminger-Finger I, Uyttendaele H, Vaudan G,

Kitajewski J, Sappino AP and Montesano R. (2000). Adv.Exp. Med. Biol., 480, 175–184.

Spahn L, Petermann R, Siligan C, Schmid JA, Aryee DN andKovar H. (2002). Cancer Res., 62, 4583–4587.

Thai TH, Du F, Tsan JT, Jin Y, Phung A, Spillman MA,Massa HF, Muller CY, Ashfaq R, Mathis JM, Miller DS,Trask BJ, Baer R and Bowcock AM. (1998). Hum. Mol.Genet., 7, 195–202.

Wu LC, Wang ZW, Tsan JT, Spillman MA, Phung A, Xu XL,Yang MC, Hwang LY, Bowcock AM and Baer R. (1996).Nat. Genet., 14, 430–440.

Xia Y, Pao GM, Chen HW, Verma IM and Hunter T. (2003).J. Biol. Chem., 278, 5255–5263.

Yu X and Baer R. (2000). J. Biol. Chem., 275, 18541–18549.

Nuclear–cytoplasmic translocation of BARD1CE Jefford et al

3520

Oncogene

Page 82: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

73

3. BARD1 expression during spermatogenesis is associated with apoptosis and hormonally regulated Feki A, Jefford CE, Durand P, Harb J, Lucas H, Krause KH, Irminger-Finger I. Biol

Reprod. 2004 Nov; 71(5):1614-24.

Since we know that BARD1 expression is high and necessary during embryogenesis

and in testis, we investigated the expression pattern of BARD1 at the mRNA and

protein levels during spermatogenesis in rat testis. In this study, we showed that

BARD1 is expressed at all stages of spermatogenesis. This fact contrasts with

BRCA1 which is only expressed in meiotic spermatocytes and early round

spermatids. While spermatogonia expressed full-length BARD1 mRNA, later stage

spermatocyte precursors express predominantly a novel, shorter splice variant

BARD1β. This BARD1β form lacks the RING finger domain, which abrogates its

interaction with BRCA1, but conserves its apoptosis-inducing capabilities.

Furthermore, it is well known that apoptosis is necessary to regulate the

overproduction of early germ cells in testis. Therefore, in order to assess the role of

BARD1 and apoptosis in testis, we monitored the expression pattern of BARD1 with

that of apoptotic markers such as BAX, active caspase-3, and DNA fragmentation on

histological sections of rat testis. We found that BARD1 expression correlates with

apoptosis as indicated by BAX and active caspase-3 coincidental spatial and

temporal expression. Furthermore, we tested and showed the ability of BARD1β

splice variant to trigger apoptosis in vitro by transient over-expression. Finally, we

show that elevated BARD1 protein expression correlates with testis pathologies

such as obstructive azoospermia, cryptorchidy and Sertoli cell syndromes. My

contribution consisted in providing the BARD1-EGFP constructs, performing the

transfections, some of the western blot as well as some technical help on apoptosis

assays are shown in Figures 3 and 5.

Page 83: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

1614

BIOLOGY OF REPRODUCTION 71, 1614–1624 (2004)Published online before print 7 July 2004.DOI 10.1095/biolreprod.104.029678

BARD1 Expression During Spermatogenesis Is Associated with Apoptosisand Hormonally Regulated1

Anis Feki,3,4 Charles-Edwards Jefford,3 Philippe Durand,5 Jean Harb,6 Herve Lucas,4 Karl-Heinz Krause,3and Irmgard Irminger-Finger2,3

Biology of Aging Laboratory,3 Department of Geriatrics, Department of Gynecology and Obstetrics,4 UniversityHospitals, CH-12225 Geneva, SwitzerlandINSERM-INRA U 418,5 Hopital Debrousse, Lyon, 69322 Cedex 05, FranceInstitut de Biologie,6 INSERM, U419 Nantes, France

ABSTRACT

The BRCA1-binding RING-finger domain protein BARD1 mayact conjointly with BRCA1 in DNA repair and in ubiquitination,but it may also induce apoptosis in a BRCA1-independent man-ner. In this study, we have investigated BARD1 expression duringspermatogenesis. In contrast with BRCA1, which is expressedonly in meiotic spermatocytes and early round spermatids,BARD1 is expressed during all stages of spermatogenesis. How-ever, while spermatogonia expressed full-length BARD1 mRNA,later stages of spermatocyte precursors express predominantlya novel, shorter splice form BARD1b. BARD1b lacks theBRCA1-interacting RING finger but maintains its proapoptoticactivity. Consistently, BRCA1 can counteract the proapoptoticactivity of full-length BARD1 but not of BARD1b. Several linesof evidence suggest that BARD1 is involved in proapoptotic sig-naling in testis: i) both BARD1 isoforms are mostly found in cellsthat stain positive for TUNEL, Bax, and activated caspase 3; ii)BARD1b, capable of inducing apoptosis even in the presence ofBRCA1, is specifically expressed in BRCA1-positive later stagesof spermatogenesis; iii) antiapoptotic hormonal stimulationleads to BARD1 downregulation; and iv) BARD1 expression isassociated with human pathologies causing sterility due to in-creased germ cell death. Our data suggest that full-lengthBARD1 might be involved in apoptotic control in spermatogoniaand primary spermatocytes, while a switch to the BRCA1-inde-pendent BARD1b might be necessary to induce apoptosis inBRCA1-expressing meiotic spermatocytes and early round sper-matids.

apoptosis, BARD1, BRCA1, FSH, infertility, meiosis, spermato-genesis, splicing, steroid hormones, testosterone

INTRODUCTION

Spermatogenesis entails a complex series of events inwhich germ cells proceed through successive rounds of mi-tosis, meiosis, and cellular differentiation and become ma-ture spermatozoa. Maturation of germ cells is closely linkedto Sertoli cells; however, their nourishing and protectivefunction is limited, and the overproduction of early germ

1Supported by focused giving grant from Johnson and Johnson and SwissNational Science Foundation grant to I.I.F.2Correspondence: Irmgard Irminger-Finger, Biology of Aging Laboratory,Department of Geriatrics, University of Geneva, 2 ch. Petit-Bel-Air, 1225Chene-Bourg/Geneva, Switzerland. FAX: 41 22 305 5455;e-mail: [email protected]

Received: 15 March 2004.First decision: 6 April 2004.Accepted: 25 June 2004.Q 2004 by the Society for the Study of Reproduction, Inc.ISSN: 0006-3363. http://www.biolreprod.org

cells is balanced by selective apoptosis [1, 2]. In the adultrat, A2–4 spermatogonia and spermatocytes undergo apo-ptosis [2, 3], thus maintaining a constant number of cellsentering meiosis. Spontaneous germ cell apoptosis can beenhanced after various insults, including testicular injuries,withdrawal of hormonal support [4, 5], radiation [6], toxi-cant [7, 8] and heat exposure [9]. Hence, whether malegerm cells survive or die is determined by a complex net-work of signals. These include paracrine signals such asstem cell factor (SCF), leukemia inhibitory factor (LIF),and desert hedgehog (Dhh), as well as endocrine signalssuch as pituitary gonadotrophs and testosterone [10]. Likeother cell types, male germ cells respond to external signalsand their internal milieu by activating intracellular signalingpathways that ultimately determine their fate [10].

In the adult rat, the rate of spontaneous apoptosis extendsto 75% and occurs mostly in spermatogonia [11–13]. Ap-optotic signaling pathways lead to caspase 3 activation andinvolve Bax and Apaf1 or the FasL/Fas and FADD [14].Expression of p53 was also reported in spermatogonia andearly meiotic prophase cells associated with DNase I ex-pression and commitment for apoptosis [15]. The upregu-lation of p53 might either lead to the repair of DNA byinducing cell cycle arrest or apoptosis by activating theproapoptotic gene Bax, a pathway well described in sper-matogenesis [16].

BARD1, a protein that interacts with p53 in vitro [17],originally identified as binding partner for BRCA1 [18], isalso implicated in the apoptotic pathway involving p53[17]. The expression of BARD1 is highly elevated in testis[19, 20], suggesting a role of BARD1 in spermatogenesis.

The mRNAs of BRCA1 and BARD1 are quite abundantin preparations from whole testis [19, 20], and an additional2 kilobase (kb) smaller BARD1 transcript of 2 kb is foundin testis, but not in other tissues, by Northern blotting. Mea-surements of BRCA1 mRNA levels in purified somaticcells of the testis and in staged germ cells showed that high-level BRCA1 mRNA expression is limited to the germ cell.Within the germ cell lineage, the high expression was de-tected in meiotic cells, specifically pachytene spermatocytesand in postmeiotic round spermatids. In contrast, little orno BRCA1 mRNA was found in premeiotic germ cells[21]. In situ hybridization localized BRCA1 mRNA to earlymeiotic prophase spermatocytes [22]. A partial understand-ing of BRCA1 function in meiosis comes from studies onmeiotic chromosomes where BRCA1 localizes to recom-bination nodules [23] and the finding that Brca1 2/2 micelack crossover during the pachytene stage and fail to enterdiplotene [24].

Page 84: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

1615BARD1 IN SPERMATOGENESIS

Although exhaustive studies on the localization andfunction of BARD1 and BRCA1 proteins during spermato-genesis are not available, portions of the BRCA1 andBARD1 proteins localize to the XIST RNA on the inactiveX-chromosome during the pachytene stage [25] and suggesta function in the maintenance of its repressed state. In thispresent study, we intend to elucidate the function ofBARD1 in spermatogenesis, including its potential role inapoptosis suggested by earlier studies on BARD1 andBRCA1.

MATERIALS AND METHODS

Tissue PreparationTwo adult male Wistar rats were used in each part of the study. Ani-

mals were anesthetized, the tunica albuginea of the right testis was incisedfrom the proximal to the distal pole, and the parenchyma was immersionfixed with 10% formaldehyde in 0.1 M phosphate buffer, pH 7.4. Thethoracic aorta was then cannulated and the vasculature flushed with buf-fered sodium chloride before perfusion fixation of the left and the righttestes with 2% glutaraldehyde and 2% paraformaldehyde in 0.05 M phos-phate buffer, pH 7.4. Procedures relating to the care and use of animalswere approved by the Swiss cantonal veterinary office.

ImmunohistochemistryThe immersion-fixed right and left testes were processed for paraffin

embedding. Consecutive orthogonal sections (5 mm) across the seminif-erous tubules were mounted on polylysine-coated slides. The first sectionof each set of sections was stained with hematoxylin eosin and used todetermine tubule stages. The remaining sections were further processedfor immunocytochemical staining. The endogenous peroxidase wasquenched by 15 min of incubation in 2% hydrogen peroxide in PBS.Unmasking of the epitope was carried out by boiling deparaffinized re-hydrated sections (twice for 5 min) in 10 mM citrate buffer, pH 6.0, usinga microwave oven at 5000 W power output. Primary antibodies were asfollows: anti-BARD1 N19 polyclonal antibody (1:20) directed to N-ter-minus of BARD1 (Santa Cruz); anti-BARD1 C20 polyclonal antibody (1:20), directed to carboxyl-terminus of BARD1 (Santa Cruz); anti-BARD1JH3 polyclonal antibody (1:20), directed to the amino acids 527–540; anti-BRCA1 D16 (1:50), epitope corresponding to the amino-terminus; anti-BRCA1 M20 (1:20), directed to the carboxyl terminus; anti-p53 (1:20)(Santa Cruz), epitope corresponding to amino acids 1–393. Secondary an-tibodies were peroxidase coupled and negative controls were processed inan identical manner except that the primary antibody was replaced by PBS.

Generation of Polyclonal Anti-BARD1 Antibody JH-3Peptides were generated corresponding to the region including amino

acids 527–540 and polyclonal antibodies JH-3 were generated in rabbits.Whole sera were used for Western blotting and immunohistochemistry.

Reverse Transcription-Polymerase Chain ReactionTotal RNA was purified from physically isolated rat germ cells of

different stages using Qiagen reagent. RNA was reverse transcribed usingoligo-dT and superscript II (Life Technologies). BARD1 primers usedwere: 56 forward (ccatggaaccagctaccg) and 2307 reverse (tcagctgtcaagag-gaagca). GAPDH primers were 176 forward (CTACCCACGGCA-AGTTCAAT) and 557 reverse (ACTGTGGTCATGAGCCCTTC). Re-verse transcription (RT) products were subjected to PCR analysis usingspecific primers for rat BARD1 cDNA (covering BARD1 coding regions).PCR cycles were: 948C 15 min, 948C 1 min, 568C 1 min, 728C 2 min 30(35 3), then 728C 10 min. Equal volumes of PCR products were analyzedon 1% agarose gel.

Western BlotBARD1 expression was monitored by Western blot analysis. Germ

cells harvested by elutriation were lysed in SDS loading buffer (4% SDS,10% glycerol, 4% 2-mercaptoethanol, 0.125 M Tris base, and 0.02% brom-ophenol blue, pH 6.8) containing protease inhibitors. Protein concentra-tions were determined using Bio-Rad protein analysis solution (Bio-Rad,Hercules, CA). Protein samples (40 mg) were size separated on 10% dis-

continuous polyacrylamide gels and transferred overnight to polyvinyli-dene fluoride membranes. Filters were blocked for 1 h with 5% nonfatmilk and then incubated for 2 h with polyclonal BARD1 antibodies (C20:1:200; JH3:1:1000). N-19, which was used in immunohistochemistry, didnot work on Western blots. After washing, membranes were incubatedwith horseradish-peroxidase-conjugated anti-rabbit antibodies for JH3 con-jugated anti-goat antibodies for C20 (1:1000), and the resulting complexeswere visualized by enhanced chemiluminescence autoradiography.

Preparation of Sertoli Cell Conditioned MediaSertoli cells were plated in tissue culture flasks at a density of 2 3 105

cells/cm2 and were cultured for 3 days in HEPES-buffered F12/DMEMsupplemented with insulin (10 mg/ml), transferrin (10 mg/ml), vitamin C(1024 M), vitamin E (10 mg/ml), retinoic acid (3.3 3 1027 M), retinol (3.33 10–7 M), pyruvate (1 mM) (all products from Sigma, La Verpilliere,France), 0.2% fetal calf serum (Life Technologies, Cergy-Pontoise, France)in the absence or presence of 1027 M testosterone (Sigma) and/or 1 ng/ml bovine NIH FSH-20 obtained through NIDDK (lot no. AFP-7028D).On Day 3, media were replaced by serum-free media supplemented or notwith hormones. On Day 5, culture media were recovered, centrifuged toeliminate cell debris, and conserved at 2208C until used. Four Sertoli cellconditioned media (SCCM) were prepared: control (no FSH, no testoster-one [F2/T2]), FSH (1 ng/ml) (F1/T2), testosterone 1027 M (F2/T1),and a combination of both hormones (F1/T1).

Isolation of Rat Sertoli Cells, Pachytene Spermatocytes,Round Spermatids, and Spermatogonia/PreleptoteneSpermatocyte Fractions

Sertoli cells from 20-day-old rats and pachytene spermatocytes andround spermatids from 90- to 120-day-old rats were isolated as describedpreviously [26]; the purity of these different fractions was assessed byflow cytometry. In Sertoli cell fractions, 77% 6 1% of the isolated cellswere 2C cells, 5.5% 6 1% were 4C cells; in pachytene spermatocytefractions, 94% 6 3% of cells were 4C cells, 3% 6 2% were 2C cells, and1% 6 0.5% were 1C cells; in round spermatid fractions, 67.2% 6 5.2%were 1C cells, 10.6% 6 6.6% were 4C cells, and 9.1% 6 1.6% were 2Ccells. Procedures relating to the care and use of animals were approvedby the French Ministry of Agriculture according to the French regulationsfor animal experimentation.

Spermatogonia/preleptotene spermatocytes were prepared by centrifu-gal elutriation of germ cells isolated from 21- to 23-day-old rats, whichdo not yet produce round spermatids in their testes. These spermatogonia/preleptotene spermatocytes eluted in the fraction of elutriation in whichround spermatids from mature rats are recovered. In this fraction, 64.8%6 7.3% of cells were 2C cells, 29.1% 6 5.4% were 4C cells, and 5.3%6 2.8% were 1C cells.

Incubation of Germ Cells in Sertoli Cells ConditionedMedia

Pachytene spermatocytes, round spermatids, spermatogonia/prelepto-tene spermatocytes were incubated for 24 h in the different conditionedmedia. At the end of the incubation period, the cells were centrifuged andrecovered for RNA extraction.

Generation of BARD1b Expression CloneBARD1b expression clone was produced by cloning the sequence of

BARD1b PCR product cDNA into pcDNA3-EGFP in frame with EGFP.Transfections were performed reproducibly with Effecten (Qiagen).

One to 4 mg of DNA was transfected per well of mouse embryonic stemcells, CGR8 [17]. Transfection efficiency was between 20% and 30% asassayed by parallel transfection of GFP cloned into pcDNA3 and analyzedby flow cytometry and fluorescent microscopy. Thirty-six hours aftertransfection, cells were harvested and apoptosis was monitored by 7-AADviability probe (BD Pharmingen) and TUNEL assay (Molecular Probes)by flow cytometry (DAKO galaxy pro).

RESULTS

BARD1 expression is most abundant in testis, as is thecase for BRCA1 [19, 20]. While the function of BRCA1had been related to meiotic recombination, based on colo-calization of BRCA1 with RAD51 to recombination nod-

Page 85: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

1616 FEKI ET AL.

FIG. 1. Immunohistology of BARD1 andBRCA1 expression in adult rat testis. A)Seminiferous tubes showing BARD1 stain-ing. BARD1 is present in spermatogonia(a) and preleptotene spermatocytes andpachytene spermatocytes (b). BARD1 anti-bodies also stain material in Sertoli cell. B)Seminiferous tubes showing BRCA1 stain-ing. BRCA1 expression is predominant atthe meiotic pachytene stage (c) and lessfrequent at the preleptotene stage (d). C)Background staining with omission of pri-mary antibody is shown.

ules [23], the function of BARD1 in spermatogenesis re-mains to be elucidated.

BARD1 Expression Is Not Correlated with BRCA1Expression

To determine the cell type and stage-specific expressionof BARD1, we performed immunohistochemistry on sec-tions of testis from adult rats. Interestingly, in the testis ofyoung adult rats (23 days) BARD1 protein is expressed inmany, but not all, spermatogonia and preleptotene sper-matocytes, and it is less abundant during later stages ofspermatogenesis (Fig. 1A). Staining was also observed insome cells at later stages of meiosis and in round sperma-tids (Fig. 1A). During early stages, BARD1 staining is lo-calized to the nucleus, but it is also found in the cytoplas-mic prolongations of Sertoli cells. The same staining couldbe observed when different antibodies against the N-ter-minal region of BARD1 were used. No staining was ob-served when the primary BARD1-specific antibody wasomitted (Fig. 1C). However, when JH-3, an antibody di-rected against the middle region of BARD1 was used, sper-matogonia showed strong staining but also later stage germcells were stained.

Because it could be expected that BARD1 and BRCA1act in concert during spermatogenesis, we determined thespecific stages at which the BRCA1 protein is expressedby immunohistochemistry with anti-BRCA1 antibodies.Consistent with previous observations of BRCA1 mRNAexpression, the BRCA1 protein was predominantly detectedat the pachytene stage and rarely in preleptotene spermato-cytes or spermatogonia (Fig. 1B). These data demonstratethat BRCA1 expression is mostly limited to the pachytenestage, while BARD1 expression is found in spermatogoniaand preleptotene spermatocytes. The elevated expression ofBARD1 but not BRCA1 in spermatogonia and primaryspermatocytes is indicative of a BRCA1-independent func-tion of BARD1 in spermatogenesis.

Stage-Specific Isoforms of BARD1

To characterize the expression pattern of BARD1 further,isolated cells from different stages of spermatogenesis wereanalyzed on Western blots. Two forms of BARD1, corre-

sponding to a molecular mass of 97 and 74 kDa, werefound on Western blots probed with different antibodies(Fig. 2A). Spermatogonia were prepared from 9-day-oldrats, at which age spermatogenesis is not induced [27], andspermatogonia but not primary spermatocytes are present.While in spermatogonia from very young rats (9 days), the97-kDa BARD1 was expressed, it was less abundant inpreleptotene spermatocytes from young rats (23 days). The74-kDa protein was found in pachytene spermatocytes ofyoung (23 days) and adult rats (3 mo), as well as in roundspermatids. Depending on which antibody was used for de-tection, the 97-kDa and 74-kDa proteins could be detectedat the preleptotene stage. This protein expression patternsuggests that a shorter isoform of BARD1 is expressed atall stages with the exception of spermatogonia. The 97-kDaBARD1 was also found in Sertoli cell preparations butmight be derived from BARD1 staining of residual bodiesphagocytosed by the Sertoli cells (data not shown). Simi-larly, p53 and Bax immunohistochemical staining onphagocytosed residual apoptotic bodies has been reportedpreviously [15, 28].

Stage-Specific BARD1 mRNA Expression

To determine whether stage-specific isoforms of BARD1were generated at the protein or the mRNA level, we in-vestigated BARD1 mRNA expression at different stages ofspermatogenesis. For this purpose, rat testes were dissectedand differentially-staged germ cells were fractionated [26].Total RNA was prepared from each of the isolated celltypes and RT-PCR was performed to monitor BARD1mRNA expression profiles. Primers were designed for am-plification of the entire BARD1 coding region. While con-trol RT-PCR of the GAPDH shows similar expression lev-els in all cell types tested, BARD1 mRNA was found inspermatogonia and less abundantly in primary spermato-cytes and round spermatids. Interestingly, RT-PCR pro-duced two products of slightly different size (Fig. 2B). Theprimers corresponding to regions of the translation initia-tion and stop codon were expected to amplify a 2251-basepair (bp) fragment, corresponding to the coding region ofwild-type BARD1. This was only found in spermatogoniaand was less pronounced in preleptotene stages. A fragment

Page 86: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

1617BARD1 IN SPERMATOGENESIS

FIG. 2. Stage-specific expression of BARD1. A) Western blots performedon cell extracts from purified stages of germ cells: spermatogonia from 9-day-old rats (G), preleptotene spermatocytes from 23-day-old rats (PL),pachytene spermatocytes of 23-day-old rats (Pjr), pachytene spermato-cytes of 3-mo-old rats (P), and round spermatids from 3-mo-old rats (RS).Two proteins of 97 and 74 kDa can be detected with antibodies JH-3 andC-20, corresponding to the full-length BARD1 protein, 97 kDa, and thecalculated molecular mass of the testis-specific BARD1b, which is about74 kDa. B) The mRNA expression in different stages of spermatogenesis.RT-PCR was performed in isolated germ cells as described in (A). Expectedlength of 2251 bp was amplified in G; two bands, 2251 bp and 1964 bp,were amplified in samples from PL stage, and 1964 bp in all other stages.Control primers were directed against the coding region of housekeepinggene GAPDH.

of approximately 2000 bp was amplified in preleptoteneand pachytene spermatocytes as well as in round spermatids(Fig. 2B). As observed on the protein level, on the tran-script level, two different mRNAs are transcribed in astage-specific fashion, full-length BARD1 in spermatogoniaand a shorter form in all stages except spermatogonia. Inpreleptotene stages, both BARD1 mRNAs and both pro-teins of 97 and 74 kDa are expressed. This might reflectthe purification procedure, which does not permit the sep-aration of spermatogonia and preleptotene spermatocytes.

A Testis-Specific Isoform of BARD1 Is Generatedby Differential Splicing

We cloned and sequenced the RT-PCR fragments cor-responding to the 2.2-kb and the approximately 2000-bpfragments. The 2.2-kb fragment was identified as 2251 bpof full-length rat BARD1 cDNA [29]. The 2000-bp frag-ment encoded a 1964-bp sequence identical to the knownrat BARD1 sequence but missing the regions including nu-cleotides 143–348 and nucleotides 1287–1367. The deletedsequences are homologous to exons 2 and 3, and 5, re-spectively, of the corresponding human BARD1 sequence(Fig. 3A). To determine whether the equivalent testis-spe-cific isoform, called BARD1b hereafter, is unique to therat, RT-PCR was confirmed on RNA isolated from mousetestis. Cloning and sequencing RT-PCR products showedthat a similar splice variant exists in the mouse. The abun-dance of mouse BARD1b when cloned from total RNAfrom mouse testis was fourfold lower that full-lengthBARD1 cDNA. Our data suggest that intron-exon bound-aries are conserved between mouse and human sequencesand that the same intron-exon boundaries are used in hu-man, mouse, and rat.

The missing BARD1 exons 2 and 3 encode part of theRING-finger structure, and exon 5 encodes one of the an-kyrin repeats. However, deletion of nucleotides 143–348leads to a frame shift that would result in translation of 46amino acid peptide before ending with a stop codon (Fig.3A). Alternatively, ATG at position 389, 42 nucleotidesdownstream of the deletion, could serve as translation ini-tiation codon. Translation initiation at nucleotide 389 wouldresult in a protein of approximately 70 kDa, which is con-sistent with the observed 74-kDa protein detected in pre-leptotene and pachytene spermatocytes and round sperma-tids. We used a deletion construct D-RING-EGFP (nucle-otides 1–340), designed to be translated from the ATG atposition 389 and producing a truncated BARD1-GFP fu-sion protein [30] and compared its expression withBARD1b fused to GFP. Both constructs expressed a trun-cated BARD1-EGFP fusion protein of approximately 97kDa, which can be detected with antibodies directed againstGFP (Fig. 3B), indicating that ATG at position 389 is afunctional translation initiation codon in vitro. This sup-ports the assumption that wild-type BARD1 is expressed inspermatogonia and the protein product of the splice variantBARD1b, schematically presented in Figure 3C, corre-sponds to the 74-kDa protein expressed in pachytene sper-matocytes and round spermatids.

BARD1 Expression Is Correlated with Apoptosis

The stage-specific pattern of BARD1 expression duringspermatogenesis was reminiscent of the reported occur-rence of apoptosis [14]. To confirm this correlation, we per-formed TUNEL experiments and BARD1 staining on serialsections from testes of a young adult rat. TUNEL assays

primarily produced positive staining in spermatogonia andpreleptotene spermatocytes coinciding specifically withanti-BARD1 staining when using antibody N-19 (Fig. 4A).TUNEL staining was also observed at later stages of sper-matogenesis, in round spermatids, coinciding with BARD1staining. Only a few TUNEL positive cells could be de-tected at the pachytene stage. Therefore, the spatial anddevelopmental distribution of BARD1 staining coincideswith apoptosis at all stages of spermatogenesis.

To further demonstrate the correlation between BARD1expression and apoptosis induction, we also tested the coex-pression of BARD1 with other proteins involved in apo-ptosis pathways, such as increased p53, Bax, and activatedcaspase 3. Staining with BARD1 antibodies against theamino-terminus colocalized with cells positive for activatedcaspase 3, tested on adjacent sections, mostly in spermato-gonia (Fig. 4B), while expression of activated caspase 3was also observed at later stages. When using an antibodyagainst the middle region of BARD1, JH-3, staining wasobserved in spermatogonia and later stages of spermato-genesis (Fig. 4B). Similar staining was observed for theproapoptotic protein Bax. At these stages, the expressionof BARD1 was primarily cytoplasmic, which is consistentwith an apoptotic function, as reported recently [30, 31].The increase of a C-terminal but not N-terminal epitope atthe pachytene stage and postmeiotic stages is consistentwith the expression of a new splice form BARD1b at thosestages and suggests that BARD1 and BARD1b expressionis associated with apoptotic events at all stages of sper-matogenesis.

BARD1b, a Potent Apoptosis Inducer

To test whether BARD1b is capable of apoptosis induc-tion, the BARD1b sequence was cloned into pcDNA andwas transiently transfected into mouse embryonic stemcells, which were previously reported to readily induce ap-optosis in response to exogenous expression of BARD1[17]. The testis-specific splice variant BARD1b induced ap-

Page 87: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

1618 FEKI ET AL.

FIG. 3. Differentially-spliced testis-specific BARD1. A) Alignment of deduced protein sequence of BARD1b (testis) with human, mouse, and rat BARD1cDNAs. RING-finger and ankyrin motifs are boxed. Regions deleted in BARD1b are indicated. Potential translation of nucleotides 1–142 would end ina stop (*) and produce a protein of 46 amino acids. An alternative ATG initiation codon at position 389 is indicated in blue (M). In-frame deletion ofexon 5 is presented within the ankyrin repeats. B) Transient expression of BARD1-EGFP (124 kDa), DRING-EGFP (97 kDa), and BARD1b-EGFP (97kDa) and analysis on Western blots probed with anti-GFP. Protein of identical size is detected after transfection of delta-RING-EGFP, and BARD1b-EGFP. C) Schematic presentation of BARD1 and presumed BARD1b protein. BARD1 is presented with RING finger (green), ankyrin repeats (blue), andBRCT repeats (red), as well as approximate location of BARD1 antibodies. BARD1b is presented with presumed location of introns (triangles), translatedregion (filled heavy line), transcribed region (thin line), and deleted regions (dotted lines). Epitopes recognized by particular antibodies (l) are indicated.

optosis in vitro as well as the full-length BARD1 sequence,pcBARD1. Apoptosis was quantified by vital staining ofliving cells with 79AAD and analyzed by flow cytometry(Fig. 5A). Interestingly, the RING-finger deficient spliceform BARD1b is as efficient in apoptosis induction asBARD1. This apoptotic capacity could be due to the lackof the RING finger, thus liberating BARD1 from apoptosis-competing binding to BRCA1, which was found previously[17]. Indeed, cotransfection of BARD1b with BRCA1 didnot lead to a reduction of apoptosis induction, while co-transfection of BRCA1 with BARD1 decreased apoptoticcapacity of BARD1 (Fig. 5, A and B). Apoptosis inductionby BARD1b could be due to increased protein stability ofBARD1b, as compared with full-length BARD1, becausean N-terminal deletion renders BARD1 more stable [30].

Apoptosis induction by BARD1 is dependent on a func-tional p53 [17]; we therefore tested whether BARD1b ex-pression affected p53 stability. Indeed, p53 is upregulatedin cells that are transfected with BARD1 and even more soin cells that are transfected with BARD1b (Fig. 5C). Fromthese data, we concluded that BARD1b, like BARD1, in-duces apoptosis when overexpressed in vitro, through sta-

bilization of p53. It is likely that BARD1b might have asimilar function in vivo.

BARD1 Expression Is Implicated in Pathologiesof Defective Spermatogenesis

Several pathologies are known that result in infertilitydue to abortive spermatogenesis, such as obstructive azo-ospermia, Sertoli cell syndrome, and cryptorchidy. Thosediseases are marked by increased levels of apoptosis ofgerm cells at different stages of spermatogenesis. We haveanalyzed the expression of BARD1 in sections of humantestis samples of those pathologies. An upregulation ofBARD1 expression is found in round spermatids and inapoptotic germ cells phagocytosed by the Sertoli cells onsections from patients suffering from azoospermia (Fig.6A). The expression of BARD1 clearly coincides with ap-optosis, as can be determined by the formation of apoptoticbodies. Cryptorchidism in humans is known to prevent allspermatogenesis in the undescended testis and germ celldevelopment is inhibited. We investigated the expression ofBARD1 in the cryptorchid testis and the contralateral de-

Page 88: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

1619BARD1 IN SPERMATOGENESIS

FIG. 4. BARD1 expression is correlated with apoptosis. A) BARD1 staining, as observed with antibody N-19, colocalizes with TUNEL staining onadjacent section (a and b). Bar 10 5 mm. B) BARD1 staining colocalizes with staining against activated caspase 3. Regions where activated caspase 3staining was observed (a), correspond to staining pattern with N-19 (b). Magnification in insets (a) and (b) was fourfold. C) Bax staining was observedat all stages. Note that BARD1 staining with antibody JH3 is cytoplasmic in most cells. Bar 5 10 mm.

Page 89: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

1620 FEKI ET AL.

FIG. 5. Apoptosis induction by testis-spe-cific BARD1b. Transient transfection assayswere performed in ES cells CGR8. A) Theextent of apoptosis induction was mea-sured by vital staining and monitored byflow cytometry. Proportion of vital cells(R2), apoptotic cells (R3), and necroticcells (R4) were measured in CGR8, CGR8plus doxorubicin (doxo), CGR8 plusBARD1b, CGR8 plus full-length BARD1,CGR8 plus BARD1b plus BRCA1, CGR8plus full-length BARD1 plus BRCA1. B)Histogram presenting percentage of celldeath, as determined in three differenttransfection experiments. C) Increased p53protein expression after transfection withBARD1b. Western blot of cell extracts af-ter transfection experiments or doxorubicintreatment demonstrates p53 increase afterBARD1b, BARD1, and BARD1b plusBRCA1 transfection, but not after transfec-tion of BARD1 plus BRCA1. Ponseau stain-ing of the membrane was applied to con-firm loading of equal amounts of protein.

scended testis of the same individual (Fig. 6B). Within thedescended testis, heavy BARD1 staining could be ob-served, and increased apoptosis and BARD1 expressionwas clearly associated with apoptotic cells and phagocy-tosed apoptotic bodies within the Sertoli cells. This is con-sistent with the reported reduced fertility of individualswith cryptorchidy. In the cryptorchid testis, which containsonly Sertoli cells, no BARD1 staining could be detected(Fig. 6, B, a and b), indicating that BARD1 is not expressedin Sertoli cells per se. These data are consistent with ourfinding that BARD1 mRNA is not expressed in Sertoli cellsand supports our interpretation that BARD1 antibody stain-ing in Sertoli cells is derived from the incorporation ofapoptotic germ cells.

BARD1 expression therefore is clearly associated withgerm cell apoptosis in vivo and with apoptosis in pathol-ogies marked by reduced spermatogenesis.

Regulation of BARD1 Transcription

Since BARD1 is associated with developmental andpathological events of apoptosis, it was interesting to in-vestigate the conditions that drive BARD1 expression. Wehave observed that BARD1 is expressed at all stages ofspermatogenesis, although to a different extent. We hy-pothesized that BARD1 expression is linked to the presenceor absence of hormones driving sexual maturation [32]. Inparticular, FSH and testosterone have been described suf-ficient for triggering spermatogenesis in vitro [33, 34]. Thisimplies that BARD1 expression is induced in spermatogo-

nia and primary spermatocytes through the paracrine sig-naling of the Sertoli cells. To test this hypothesis, we cul-tured isolated germ cells, specifically spermatogonia/pre-leptotene stage, pachytene, and round spermatids, either innormal culture medium, as defined previously [26], or inconditioned culture medium, obtained from Sertoli cellsthat were preincubated with hormones, to mimic in vivoparacrine signaling and maturation. FSH and testosteronewere added to the culture medium. The response of BARD1to hormone exposure, FSH or testosterone or both, wasmonitored on the mRNA level. RT-PCR was performed todetect BARD1 mRNA expression in RNA isolated from thedifferent cultures (Fig. 7).

Interestingly, in spermatogonia/preleptotene germ cells,no reduction of expression was observed with the additionof testosterone. FSH, however, led to a complete repressionof BARD1, but the combination of FSH and testosteronehad no effect, suggesting that testosterone compensated theeffect of FSH. In all conditions, two forms of BARD1mRNA were detected that can be correlated with the wild-type full-length BARD1 and with BARD1b, respectively,when primers amplifying regions including nucleotides 56–2307 were used (Fig. 7A). At the pachytene stage, the ad-dition of hormones did not result in a variation of the ex-pression level of BARD1. At the round spermatid stage,addition of hormones produced a similar repression pattern,as observed at the spermatogonia/preleptotene stage, buttestosterone had no inhibitory effect on FSH-induced re-pression of BARD1. Both forms of BARD1, full length and

Page 90: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

1621BARD1 IN SPERMATOGENESIS

FIG. 6. BARD1 expression in human pa-thologies. A) BARD1 staining shown oncross-section of seminiferous tubes. Thesyndrome of azoospermia is characterizedby abortive spermatogenesis and increasedformation of apoptotic bodies, which showBARD1 staining (arrow heads) at stagescorresponding to round spermatids. Fewgerm cells are developed in Sertoli cellsyndrome. Germ cells phagocytosed bythe Sertoli cells are stained positively forBARD1 (arrows). B) Human cryptorchidy.Hematoxylin and eosin staining showssuppression of germ cell development andpresence of only a few Sertoli cells (a),and BARD1 staining is negative (b) in theintra-abdominal testis. In the contralateraldescended testis, hematoxylin stainingshows a large number of condensed, apo-ptotic nuclei (c). BARD1 staining (d) is lo-calized to cells of different stages, presum-ably spermatogonia (arrow 1), preleptotene(arrow 2), and pachytene (arrow 3). In-creased number of apoptotic cells and ap-optotic bodies are observed in associationwith BARD1 staining. A and B) Originalmagnification 340.

BARD1b, were obtained, although BARD1b was increasedproportionally. These results indicate that FSH acts as atranscriptional repressor of BARD1 in spermatogonia/pri-mary spermatocytes and round spermatids but has no effecton pachytene stage spermatocytes. Interestingly, however,testosterone, by itself, does not have any influence onBARD1 mRNA expression at any of the stages tested, butin combination with FSH, it annihilates the repressive effectof FSH.

In contrast, when cells were grown in preconditionedmedium, medium from Sertoli cells that had been exposedto the hormones, the addition of FSH and FSH plus testos-terone, but not testosterone alone, resulted in suppressionof BARD1 expression in spermatogonia/preleptotene sper-matocytes and in round spermatids. No effect of either hor-mone was observed under these conditions in pachytenespermatocytes (Fig. 7A). Interestingly, testosterone alone,when applied through Sertoli cell-conditioned media, led toBARD1 repression in round spermatids. Furthermore, tes-tosterone had no compensatory effect on BARD1 repres-sion by FSH, but even a repressive effect by itself in roundspermatids. These data indicate that FSH, when applied toSertoli cells, could act as a messenger for BARD1 repres-sion at the spermatogonia/preleptotene stage and in roundspermatids.

To test whether other genes, the products of which pre-sumably interact with BARD1, were also hormonally reg-ulated in this in vitro system, we monitored the expressionof p53 mRNAs (Fig. 7B). The expression of p53 showedonly minor, insignificant changes under conditions of FSHand/or testosterone addition. No changes were observed for

the control gene b-tubulin, which remained unchanged inthe various conditions.

In summary, our data suggest that, at the preleptoteneand round spermatid stages, BARD1 is directly or indirectlyrepressed by FSH through paracrine signaling of the Sertolicells. The same treatment affects only mild repression ofBRCA1 and only in preleptotene cells (data not shown).

DISCUSSION

Two mechanisms seem to be of crucial importance dur-ing spermatogenesis: meiotic recombination and apoptosis.BARD1 and BRCA1 were reported to have maximal ex-pression in testis. BRCA1 is mainly expressed during mei-osis [21–24], where, most likely, it is involved in meioticrepair. BARD1, however, is mainly expressed in spermato-gonia and primary spermatocytes and to a lesser extent dur-ing and after meiosis. BARD1 expression rather correlateswith apoptosis, which is known to occur at several stagesof spermatogenesis, including spermatogonia [14]. BARD1specifically colocalized with TUNEL staining (Fig. 4A) andwith proteins implicated in apoptotic pathways, such as ac-tivated caspase 3, p53, and Bax (Fig. 4B). However, thereare more cells expressing BARD1 than apoptosis effectorgenes, such as Bax. This could reflect the fact that BARD1acts upstream of the apoptosis pathway and that other ad-ditional signals are required to activate apoptosis. Thesedata are consistent with a role of BARD1 in a caspase-dependent apoptotic pathway, as reported previously [17].Other functions cannot be ruled out, however, such as thebinding of BARD1 together with BRCA1 to XIST RNA

Page 91: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

1622 FEKI ET AL.

FIG. 7. Hormonal regulation of BARD1 expression during spermatogen-esis. Purified germ cells, spermatogonia/preleptotene (GPL), pachytene(P), or round spermatids (RS) were cultured. A) BARD1 expression afterdirect addition of FSH (F) and/or testosterone (T) or after incubation withSertoli cell-conditioned medium (SCCM), was monitored by RT-PCR.SCCM was prepared as described in the Materials and Methods. Bothforms of BARD1 (2.2 kb and 2.0 kb) are coexpressed. A third and muchless abundant band can be observed and represents a form without de-letion of exon 5 but not exon 2 and 3. Control RT-PCR was performedwith primers directed against b-tubulin. B) The p53 or b-tubulin (b-tub)expression of cultured germ cells was monitored after incubation withSCCM medium. Sertoli cells were cultured with FSH and/or testosteronefor 1 day and culture medium was added to germ cell cultures. RT-PCRwas performed to compare the expression profiles of BARD1, p53, andthe housekeeping gene b-tubulin.

on the inactive X chromosome [25]. BARD1 interactionwith BRCA1 might compete for binding to other proteinsand inhibit its apoptotic activity [17]. Consequently, theexpression of BARD1 in the absence of BRCA1 could leadto increased apoptotic capacity.

BARD1 knockouts have been generated but are lethalduring the early stages of embryogenesis and thus cannotshed light on the function of BARD1 during spermatogen-esis. We therefore performed experiments to further dissectthe specific local and temporal expression pattern ofBARD1 on the mRNA and the protein level. We found thata novel isoform of BARD1, generated by differential splic-ing, was expressed in the testis in addition to wild-typeBARD1. The testis-specific new splice variant, BARD1b,shows a deletion of exons 2, 3, and 5. The translation ofBARD1b results in a frame shift after excision of exons 2and 3 and would end with a stop codon (Fig. 3A). However,downstream of the splice acceptor site, an ATG at position389 could act as an alternative translation initiation site.Western blot analysis is consistent with this site of trans-lation initiation because the proteins detected with anti-BARD1 antibodies correspond to the expected size for thepresumed translation product of BARD1b. Expression of adeletion construct, missing regions 1–380, results in a pro-tein of similar size as translation of BARD1b. It is thereforelikely that the 74-kDa protein, expressed at all stages ofspermatogenesis with the exception of spermatogonia anddetected with antibodies against the C-terminal portion of

BARD1, is a translation product of the differentially splicedform BARD1b.

The temporal expression patterns of BARD1 mRNAsand BARD1 proteins isoforms show a similar distribution,a 97-kDa protein corresponds to full-length BARD1 and isexpressed in spermatogonia, as is the full-length BARD1transcript. BARD1b mRNA is expressed at all stages ex-cept in spermatogonia, as is a 74-kDa protein, which isrecognized by C-terminal antibodies, suggesting that the74-kDa is derived from differential splicing rather thanfrom posttranslational modification of BARD1. Interesting-ly, the presumed protein product of BARD1b lacks theRING-finger domain, required for the interaction ofBARD1 with BRCA1 [18, 35], but retains regions sufficientfor its apoptotic function [30].

Expression of BARD1b in vitro resulted in similar levelsof apoptosis as BARD1. However, the in vivo apoptosisinduced by BARD1b could be more efficient due to in-creased stability of BARD1b compared with full-lengthBARD1, as determined on Western blot analysis using an-tibodies against different regions of BARD1 (data notshown) and due to the lack of BRCA1 interaction domain.BRCA1 could act as an inhibitor of BARD1-induced apo-ptosis [17]; therefore, we demonstrate that BRCA1 reducesBARD1-induced apoptosis but has no affect on the apo-ptotic capacity of BARD1b (Fig. 5, A and B). BARD1bnot only induces apoptosis but also shows p53 stabilization,as does wild-type BARD1 (Fig. 5C). Therefore, BARD1bacts in a p53-dependent apoptosis pathway. It could bespeculated that BARD1b acts as apoptosis inducer in stageswhere BRCA1 and BARD1 are coexpressed.

This hypothesis stipulates that, during premeiotic stages,in the absence of BRCA1, full-length BARD1 proteins areexpressed and presumably are involved in apoptotic func-tions. To ensure apoptosis induction by BARD1 in the pres-ence of the competitor BRCA1 [17], the testis-specificsplice variant BARD1b is expressed.

Besides the importance of apoptosis in normal sper-matogenesis, the loss of germ cell homeostasis, or dereg-ulated apoptosis, can result in infertility. In line with thesethoughts, BARD1 is associated with pathologies of reducedgerm cell maturation due to increased apoptosis. Loss orreduction of apoptosis, on the other hand, can lead to un-controlled proliferation and it could be speculated thatBARD1 plays a role as tumor suppressor in germ cell can-cers.

Male germ cell apoptosis and development is regulatedby different growth-promoting or -repressing signals, oneof the regulators being hormones. Testosterone deprivationleads to germ cell apoptosis [5] and estradiol increasesgerm cell survival [36, 37]. We speculated that BARD1expression was hormonally regulated because BARD1 mu-tations are implicated in female gynecological cancers [38,39] and its expression pattern in ovary and uterine tissuereflected a hormonal modulation [19]. Here we report hor-monally regulated transcriptional activation or repression ofBARD1. Interestingly, BARD1 transcription is repressed bythe addition of FSH or FSH and testosterone to isolatedgerm cells, but not testosterone alone (Fig. 7A). But eitherFSH or testosterone are capable of transcriptional repres-sion of BARD1 when added in the presence of Sertoli cells.Sertoli cells can convert testosterone to estradiol in thepresence of FSH, and estradiol is required for progressionof spermatogenesis [40]. Our data therefore suggest that therepressive effect on BARD1 transcription is due to estradiolrather than testosterone.

Page 92: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

1623BARD1 IN SPERMATOGENESIS

In pachytene spermatocytes, no effect is observed byeither hormones added, neither directly nor as conditionedmedium of Sertoli cells. In round spermatids, however, asimilar response is observed as in spermatogonia/prelepto-tene spermatocytes. It is important to emphasize that thepattern of expression of the two BARD1 splice forms isinfluenced by additional factors in vivo (Fig. 2), which arenot reproduced in vitro (Fig. 7). Interestingly, BRCA1 ex-pression parallels the expression of BARD1. This could bedue to similar transcriptional repression of BRCA1 or toBARD1 mRNA or protein affecting BRCA1 transcription,although the latter possibility is unlikely, considering ex-periments performed with xBARD1 and xBRCA1 [41]. Asexpected, the expression of p53 mRNA was only slightly,if at all, modulated. Our data are consistent with a modelthat predicts that BARD1 is repressed by FSH and/or es-tradiol produced in Sertoli cells but not by testosterone.BARD1 acts as mediator between paracrine signals and ap-optosis by transcriptional response to hormone levels, asincreased transcript and protein levels of BARD1 lead toincreased p53 levels and apoptosis [17]. A pathway ofFSH-induced apoptosis to caspase activation and apoptosisinduction has been described previously [34]. This wouldmean that cells in contact with Sertoli cells are more proneto be exposed to the suppressive effect of FSH even in thepresence of testosterone, consistent with survival ratherthan apoptosis and consistent with a limited support func-tion of Sertoli cells.

Although the correlation of BARD1 expression and ap-optosis in spermatogenesis is intriguing, it remains to beconsidered that components of the apoptotic machinery inmale germ cells are required for and activated during sper-matogenesis without completion of apoptosis, as demon-strated for Drosophila [28]. Whether similar mechanismsexist in mammals remains to be determined.

ACKNOWLEDGMENTS

We are grateful to M. F. Pelte, D. Chardonnens, A. Agostini, and G.Bianchi for contributing with their expertise and with critical comments,to W.-C. Leung for fruitful discussions, and to A. Caillon, C. Genet, M.H. Saw and M. Vigier for excellent technical assistance.

REFERENCES

1. Bartke A. Effects of growth hormone on male reproductive functions.J Androl 2000; 21:181–188.

2. Sinha Hikim AP, Swerdloff RS. Hormonal and genetic control of germcell apoptosis in the testis. Rev Reprod 1999; 4:38–47.

3. Allan DJ, Harmon BV, Roberts SA. Spermatogonial apoptosis hasthree morphologically recognizable phases and shows no circadianrhythm during normal spermatogenesis in the rat. Cell Prolif 1992;25:241–250.

4. Hikim AP, Wang C, Leung A, Swerdloff RS. Involvement of apoptosisin the induction of germ cell degeneration in adult rats after gonad-otropin-releasing hormone antagonist treatment. Endocrinology 1995;136:2770–2775.

5. Woolveridge I, de Boer-Brouwer M, Taylor MF, Teerds KJ, Wu FC,Morris ID. Apoptosis in the rat spermatogenic epithelium followingandrogen withdrawal: changes in apoptosis-related genes. Biol Reprod1999; 60:461–470.

6. Meistrich ML. Effects of chemotherapy and radiotherapy on sper-matogenesis. Eur Urol 1993; 23:136–141. Discussion 142.

7. Li LH, Wine RN, Chapin RE. 2-Methoxyacetic acid (MAA)-inducedspermatocyte apoptosis in human and rat testes: an in vitro compari-son. J Androl 1996; 17:538–549.

8. Ku WW, Wine RN, Chae BY, Ghanayem BI, Chapin RE. Spermato-cyte toxicity of 2-methoxyethanol (ME) in rats and guinea pigs: evi-dence for the induction of apoptosis. Toxicol Appl Pharmacol 1995;134:100–110.

9. Lue YH, Hikim AP, Swerdloff RS, Im P, Taing KS, Bui T, Leung A,

Wang C. Single exposure to heat induces stage-specific germ cell ap-optosis in rats: role of intratesticular testosterone on stage specificity.Endocrinology 1999; 140:1709–1717.

10. Print CG, Loveland KL. Germ cell suicide: new insights into apopto-sis during spermatogenesis. Bioessays 2000; 22:423–430.

11. Huckins C. The morphology and kinetics of spermatogonial degen-eration in normal adult rats: an analysis using a simplified classifica-tion of the germinal epithelium. Anat Rec 1978; 190:905–926.

12. De Rooij DG, Lok D, Huckins C. Regulation of the density of sper-matogonia in the seminiferous epithelium of the Chinese hamster: II.Differentiating spermatogonia. The morphology and kinetics of sper-matogonial degeneration in normal adult rats: an analysis using a sim-plified classification of the germinal epithelium. Anat Rec 1987; 217:131–136.

13. Oakberg EF. Duration of spermatogenesis in the mouse. Nature 1957;180:1137–1138.

14. Sinha Hikim AP, Lue Y, Diaz-Romero M, Yen PH, Wang C, SwerdloffRS. Deciphering the pathways of germ cell apoptosis in the testis. JSteroid Biochem Mol Biol 2003; 85:175–182.

15. Stephan H, Polzar B, Rauch F, Zanotti S, Ulke C, Mannherz HG.Distribution of deoxyribonuclease I (DNase I) and p53 in rat testisand their correlation with apoptosis. Histochem Cell Biol 1996; 106:383–393.

16. Yan W, Samson M, Jegou B, Toppari J. Bcl-w forms complexes withBax and Bak, and elevated ratios of Bax/Bcl-w and Bak/Bcl-w cor-respond to spermatogonial and spermatocyte apoptosis in the testis.Mol Endocrinol 2000; 14:682–699.

17. Irminger-Finger I, Leung WC, Li J, Dubois-Dauphin M, Harb J, FekiA, Jefford CE, Soriano JV, Jaconi M, Montesano R, Krause KH. Iden-tification of BARD1 as mediator between proapoptotic stress and p53-dependent apoptosis. Mol Cell 2001; 8:1255–1266.

18. Wu LC, Wang ZW, Tsan JT, Spillman MA, Phung A, Xu XL, YangMC, Hwang LY, Bowcock AM, Baer R. Identification of a RINGprotein that can interact in vivo with the BRCA1 gene product. NatGenet 1996; 14:430–440.

19. Irminger-Finger I, Soriano JV, Vaudan G, Montesano R, Sappino AP.In vitro repression of BRCA1-associated RING domain gene,BARD1, induces phenotypic changes in mammary epithelial cells. JCell Biol 1998; 143:1329–1339.

20. Ayi TC, Tsan JT, Hwang LY, Bowcock AM, Baer R. Conservation offunction and primary structure in the BRCA1-associated RING do-main (BARD1) protein. Oncogene 1998; 17:2143–2148.

21. Zabludoff SD, Wright WW, Harshman K, Wold BJ. BRCA1 mRNAis expressed highly during meiosis and spermiogenesis but not duringmitosis of male germ cells. Oncogene 1996; 13:649–653.

22. Blackshear PE, Goldsworthy SM, Foley JF, McAllister KA, BennettLM, Collins NK, Bunch DO, Brown P, Wiseman RW, Davis BJ.BRCA1 and BRCA2 expression patterns in mitotic and meiotic cellsof mice. Oncogene 1998; 16:61–68.

23. Scully R, Chen J, Plug A, Xiao Y, Weaver D, Feunteun J, Ashley T,Livingston DM. Association of BRCA1 with Rad51 in mitotic andmeiotic cells. Cell 1997; 88:265–275.

24. Xu X, Aprelikova O, Moens P, Deng CX, Furth PA. Impaired meioticDNA-damage repair and lack of crossing-over during spermatogenesisin BRCA1 full-length isoform deficient mice. Development 2003;130:2001–2012.

25. Ganesan S, Silver DP, Greenberg RA, Avni D, Drapkin R, Miron A,Mok SC, Randrianarison V, Brodie S, Salstrom J, Rasmussen TP,Klimke A, Marrese C, Marahrens Y, Deng CX, Feunteun J, LivingstonDM. BRCA1 supports XIST RNA concentration on the inactive Xchromosome. Cell 2002; 111:393–405.

26. Weiss M, Vigier M, Hue D, Perrard-Sapori MH, Marret C, Avallet O,Durand P. Pre- and postmeiotic expression of male germ cell-specificgenes throughout 2-week cocultures of rat germinal and Sertoli cells.Biol Reprod 1997; 57:68–76.

27. Jahnukainen K, Chrysis D, Hou M, Parvinen M, Eksborg S, SoderO. Increased apoptosis occurring during the first wave of spermato-genesis is stage-specific and primarily affects mid-pachytene sper-matocytes in the rat testis. Biol Reprod 2004; 70:290–296.

28. Arama E, Agapite J, Steller H. Caspase activity and a specific cyto-chrome C are required for sperm differentiation in Drosophila. DevCell 2003; 4:687–697.

29. Gautier F, Irminger-Finger I, Gregoire M, Meflah K, Harb J. Identi-fication of an apoptotic cleavage product of BARD1 as an autoanti-gen: a potential factor in the antitumoral response mediated by apo-ptotic bodies. Cancer Res 2000; 60:6895–6900.

30. Jefford CE, Feki A, Harb J, Krause KH, Irminger-Finger I. Nuclear-

Page 93: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

1624 FEKI ET AL.

cytoplasmic translocation of BARD1 is linked to its apoptotic activity.Oncogene 2004; 23:3509–3520.

31. Rodriguez JA, Schuchner S, Au WW, Fabbro M, Henderson BR. Nu-clear-cytoplasmic shuttling of BARD1 contributes to its proapoptoticactivity and is regulated by dimerization with BRCA1. Oncogene2004; 23:1809–1820.

32. Billig H, Furuta I, Rivier C, Tapanainen J, Parvinen M, Hsueh AJ.Apoptosis in testis germ cells: developmental changes in gonadotropindependence and localization to selective tubule stages. Endocrinology1995; 136:5–12.

33. Tesarik J, Mendoza C, Greco E. The effect of FSH on male germ cellsurvival and differentiation in vitro is mimicked by pentoxifylline butnot insulin. Mol Hum Reprod 2000; 6:877–881.

34. Tesarik J, Nagy P, Abdelmassih R, Greco E, Mendoza C. Pharmaco-logical concentrations of follicle-stimulating hormone and testosteroneimprove the efficacy of in vitro germ cell differentiation in men withmaturation arrest. Fertil Steril 2002; 77:245–251.

35. Morris JR, Keep NH, Solomon E. Identification of residues requiredfor the interaction of BARD1 with BRCA1. J Biol Chem 2002; 277:9382–9386.

36. Atanassova N, McKinnell C, Turner KJ, Walker M, Fisher JS, MorleyM, Millar MR, Groome NP, Sharpe RM. Comparative effects of neo-natal exposure of male rats to potent and weak (environmental) estro-

gens on spermatogenesis at puberty and the relationship to adult testissize and fertility: evidence for stimulatory effects of low estrogenlevels. Endocrinology 2000; 141:3898–3907.

37. Pentikainen V, Erkkila K, Suomalainen L, Parvinen M, Dunkel L.Estradiol acts as a germ cell survival factor in the human testis invitro. J Clin Endocrinol Metab 2000; 85:2057–2067.

38. Thai TH, Du F, Tsan JT, Jin Y, Phung A, Spillman MA, Massa HF,Muller CY, Ashfaq R, Mathis JM, Miller DS, Trask BJ, Baer R, Bow-cock AM. Mutations in the BRCA1-associated RING domain(BARD1) gene in primary breast, ovarian and uterine cancers. HumMol Genet 1998; 7:195–202.

39. Ghimenti C, Sensi E, Presciuttini S, Brunetti IM, Conte P, BevilacquaG, Caligo MA. Germline mutations of the BRCA1-associated ringdomain (BARD1) gene in breast and breast/ovarian families negativefor BRCA1 and BRCA2 alterations. Genes Chromosomes Cancer2002; 33:235–242.

40. Robertson KM, O’Donnell L, Jones ME, Meachem SJ, Boon WC,Fisher CR, Graves KH, McLachlan RI, Simpson ER. Impairment ofspermatogenesis in mice lacking a functional aromatase (cyp 19) gene.Proc Natl Acad Sci U S A 1999; 96:7986–7991.

41. Joukov V, Chen J, Fox EA, Green JB, Livingston DM. Functionalcommunication between endogenous BRCA1 and its partner, BARD1,during Xenopus laevis development. Proc Natl Acad Sci U S A 2001;98:12078–12083.

Page 94: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

85

4. BARD1 induces apoptosis by catalysing phosphorylation of p53 by DNA damage response kinase Feki A, Jefford CE, Berardi P, Wu JY, Cartier L, Krause KH, Irminger-Finger I.

Oncogene. 2005 Mar 14. (In press).

It was previously shown that BARD1 plays a role in mediating apoptosis,

independently of BRCA1, but it requires p53 for signalling towards apoptosis. This

was later confirmed because BARD1-repressed cells exhibit resistance in signalling

apoptosis. Furthermore, p53 stability and activation requires its phosphorylation on

serine 15. In order to investigate further the interactions between BARD1 and p53,

and map the region imputable to this binding ability, we performed co-

immunoprecipitation assays and western blots. In this study we reveal that BARD1

over-expression may be sufficient to activate p53 phosphorylation on serine 15. We

identify the region of BARD1 which binds to p53 through a series of co-

immunoprecipitations of deletion mutants fused to EGFP. We can see that even the

deletion mutants lacking the RING functional domain, which normally binds BARD1

to BRCA1, are capable of co-immunoprecipitating p53. This results rules out the

preconception that p53 was co-immunoprecipitated through interaction with BRCA1.

We show also that BARD1 binds to Ku-70 in co-immunoprecipitation experiments as

well as phosphorylated p53. These result further points to a role for Ku-70 in

phosphorylating p53. My involvement consisted in furnishing the various plasmids

coding for the BARD1-EGFP fusion proteins, immunocytochemistry, co-

immunoprecipitation and western blots appearing in Figures 2 and 3. I also furnished

some of the data shown in Figure 4 by repeating the experiments in order to confirm

validity of our results.

Page 95: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

ORIGINAL PAPER

BARD1 induces apoptosis by catalysing phosphorylation of p53 by

DNA-damage response kinase

Anis Feki1,2, Charles Edward Jefford1, Philip Berardi1,3, Jian-Yu Wu1, Laetitia Cartier1,Karl-Heinz Krause1 and Irmgard Irminger-Finger*,1

1Biology of Aging Laboratory, Department of Geriatrics, University of Geneva, Chemin de Petit Bel Air 2, CH-1225Geneva/Chene-Bourg, Switzerland; 2Department of Gynecology and Obstetrics, University Hospital of Geneva, Geneva,Switzerland; 3Biochemistry and Molecular Biology and Oncology, University of Calgary, Calgary, Canada

The BRCA1-associated RING domain protein BARD1acts with BRCA1 in double-strand break repair andubiquitination. BARD1 plays a role as mediator of apop-tosis by binding to and stabilizing p53, and BARD1-repressed cells are resistant to apoptosis. We thereforeinvestigated the mechanism by which BARD1 induces p53stability and apoptosis. The apoptotic activity of p53 isregulated by phosphorylation. We demonstrate thatBARD1 binds to unphosphorylated and serine-15 phos-phorylated forms of p53 in several cell types and that theregion required for binding comprises the region sufficientfor apoptosis induction. In addition, BARD1 binds to Ku-70, the regulatory subunit of DNA-PK, suggesting thatthe mechanism of p53-induced apoptosis requires BARD1for the phosphorylation of p53. Upregulation of BARD1alone is sufficient for stabilization of p53 and phosphor-ylation on serine-15, as shown in nonmalignant epithelialcells and ovarian cancer cells, NuTu-19, which aredefective in apoptosis induction and express aberrantsplice variants of BARD1. Stabilization and phosphoryla-tion of p53 in NuTu-19 cells, as well as apoptosis, can beinduced by the exogenous expression of wild-typeBARD1, suggesting that BARD1, by binding to thekinase and its substrate, catalyses p53 phosphorylation.Oncogene advance online publication, 14 March 2005;doi:10.1038/sj.onc.1208491

Keywords: apoptosis; BARD1; DNA-PK; Ku-70; p53;phosphorylation

Introduction

Apoptosis is an important means of eliminatingdangerously damaged cells. Cancer cells have developedstrategies to escape apoptosis and consequently becomeresistant to apoptosis inducing treatments. Along this

line, the tumor suppressor p53, a key element in theapoptosis response pathway, is mutated in more than50% of tumors. Several signaling pathways converge onthe activation of p53, which leads either to cell cyclearrest and DNA repair or senescence, or apoptosis(Levine, 1997). P53 activation and stabilization dependon its phosphorylation at multiple sites, by a number ofkinases (Meek, 1994; Milczarek et al., 1997). Phosphor-ylation at serines 6 and 9 by casein kinase 1-delta andcasein kinase 1-epsilon occurs both in vitro and invivo (Knippschild et al., 1997; Kohn, 1999), thecheckpoint kinases Chk1 and Chk2 phosphorylation atserine 20 enhances its tetramerization, stability, andactivity (Shieh et al., 1999; Hirao et al., 2000).Phosphorylation of p53 at serine-392 is observed invitro by CDK-activating kinase (CAK) (Lu et al., 1997)and in vivo (Hao et al., 1996; Lu et al., 1997), but isaltered in human tumors. Serine-392 phosphorylationhas been reported to influence the growth suppressorfunction, DNA binding, and transcription activation ofp53 (Hao et al., 1996; Lohrum and Scheidtmann, 1996;Lu et al., 1997; Kohn, 1999). Phosphorylation of p53 atserine-46 is important for regulating its ability to induceapoptosis (Oda et al., 2000). However, the rapidresponse to DNA damage is induced by phosphoryla-tion of p53 at serines 15, 20, and 37 and leads to reducedinteraction of p53 with its negative regulator, oncopro-tein MDM2 (Shieh et al., 1997).Known kinases for the DNA damage response are

ATM, ATR, and DNA-PK, which can phosphorylatep53 at serines 15 and 37 (Shieh et al., 1997; Tibbettset al., 1999). These phosphorylation events impair theability of MDM2 to bind p53, promoting both theaccumulation and functional activation of p53 inresponse to DNA damage (Shieh et al., 1997; Tibbettset al., 1999). MDM2 acts as inhibitor of p53 function bytargeting it for ubiquitination and proteasomal degra-dation (Honda et al., 1997; Chehab et al., 1999).The tumor suppressor BRCA1 Associated Ring

Domain protein (BARD1) can act as a mediatorbetween genotoxic stress and apoptosis by binding toand stabilizing p53 (Irminger-Finger et al., 2001;Irminger-Finger and Leung, 2002). BARD1 is the majorprotein-binding partner of BRCA1 (Wu et al., 1996)

Received 22 July 2004; revised 8 November 2004; accepted 20 December2004

*Correspondence: I Irminger-Finger;E-mail: [email protected];URL: http://www.medecine.unige.ch/recherche/biologie_du_vieillissement/

Oncogene (2005), 1–11& 2005 Nature Publishing Group All rights reserved 0950-9232/05 $30.00

www.nature.com/onc

Page 96: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

and the BARD1/BRCA1 heterodimer has functions inDNA repair (Jin et al., 1997; Scully et al., 1997) andubiquitination (Baer and Ludwig, 2002). Both proteinsare homologous within their RING domains and twoBRCT domains. The BRCT domains of BRCA1 haverecently been characterized as modules for kinase andtarget interaction (Manke et al., 2003; Yu et al., 2003)and a role of BARD1–BRCA1 complexes in signalingfrom DNA damage to p53 phosphorylation has beendescribed recently (Fabbro et al., 2004).The repression of BARD1 expression in vitro results

in genomic instability and cellular changes suggestiveof a premalignant phenotype (Irminger-Finger et al.,1998) and resistance to apoptosis (Irminger-Finger et al.,2001). Similarly, BARD1, as well as BRCA1, knockoutcells exhibit high levels of genomic instability (Joukovet al., 2001; Ludwig et al., 2001; McCarthy et al., 2003).However, excess of BARD1 over BRCA1 results inapoptosis independently of BRCA1 (Irminger-Fingeret al., 2001). Interestingly, the expression of BARD1,but not BRCA1, is hormonally controlled in uterinetissue (Irminger-Finger et al., 1998), suggesting a rolein tissue homeostasis. Further evidence for a role ofBARD1 in tumor suppression is presented by BARD1mutations found in breast, ovarian, and most intrigu-ingly in endometrial tumors (Thai et al., 1998; Hashi-zume et al., 2001; Ghimenti et al., 2002; Ishitobi et al.,2003). In contrast, BRCA1 mutations have not beenfound in uterine cancers, supporting a BRCA1-independent role of BARD1 in tumor suppressionby apoptosis induction (Irminger-Finger and Leung,2002).To further elucidate the mechanism by which BARD1

induces apoptosis, we investigated its possible functionin the regulation of p53 phosphorylation required forapoptosis (Lakin and Jackson, 1999).Here we report that BARD1 is involved in p53ser15

phosphorylation and BARD1 overexpression is suffi-cient for this phosphorylation. We identified a cancer cellline which is resistant to apoptosis and deficient inp53ser15 phosphorylation in response to genotoxic stress.Indeed, BARD1 is mutated and deleted in this cell lineand exogenous expression can restore p53ser15 phosphory-lation and apoptosis.

Results

Given the role of BARD1 in p53 stabilization, weinvestigated its expression in relation to p53 in variouscell types. In nonmalignant mammary gland cells, TAC-2, treatment with the apoptosis-inducing drug doxo-rubicin resulted in the upregulation of BARD1 andaccumulation of the p53 protein, as shown previously(Irminger-Finger et al., 2001). Simultaneously to theincrease of BARD1 and p53 protein levels upongenotoxic stress, the accumulation of phosphorylatedp53ser15 was observed when phospho-epitope-specificantibodies against phospho-p53ser15 were used on Wes-tern blots (Figure 1a). This is consistent with the

reported stabilization of p53 by phosphorylation uponstress induction (Shieh et al., 1997). Phosphorylation ofserines 15, 20, and 37 is essential for p53 function duringapoptosis (Shieh et al., 1997; Tibbetts et al., 1999). Sinceserine 15, but not serines 20 and 37, is conservedbetween human, mouse, and rat p53, we determinedwhether this site was implicated in the resistance toapoptosis in cancer cells.

Figure 1 p53 and phospho-p53 expression correlated withBARD1 expression. (a) Doxorubicin treatment of TAC-2 cellsinduces BARD1, p53 and phospho-p53 (p-p53). Western blots wereperformed with respective antibodies on cell extracts treated withdoxorubicin for various intervals. (b) BARD1 is expressed in TAC-2 but not NuTu-19 cells after treatment with doxorubicin.Correlated expression of p53 and phospho-p53 is shown. (c)Exogenous expression of BARD1 restores p53 expression andphosphorylation. HEK 293T cells with constitutive expression ofp53 and NuTu-19 cells were transduced with lentiviral vectorexpressing BARD1. The membranes were probed with anti-actin asloading control

BARD1 induces apoptosisA Feki et al

2

Oncogene

Page 97: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

We first compared the effect of doxorubicin treatmentapplied to TAC-2 cells and to highly malignant ratovarian cancer cells, NuTu-19 (Rose et al., 1996).NuTu-19 lack the full-length BARD1 protein andshowed only a slight upregulation of p53, as comparedto TAC-2 cells (Figure 1b). Most importantly, no traceof phospho-p53 could be detected in doxorubicin-treated NuTu-19 cells, while phospho-p53 was readilydetectable in TAC-2 cells.Since overexpression of BARD1 by itself can lead to

p53 stabilization and apoptosis (Irminger-Finger et al.,2001), we tested the effect of BARD1 expression onp53 phosphorylation. HEK 293T cells were used andcompared to malignant NuTu-19 cells, since both celltypes showed similar transduction efficiency withlentiviral vectors. Transduction of HEK 293T andNuTu-19 cells with wild-type BARD1, under the controlof the EF1-alpha promotor, resulted in increasedexpression of BARD1 in HEK 293T cells and expressionof exogenous BARD1 in NuTu-19 cells. An increase oftotal p53 in HEK 293T cells, and less pronounced inNuTu-19 cells, was observed. Phosphorylation of p53p53ser15 was observed in HEK 293T cells and NuTu-19cells after exogenous expression of BARD1 (Figure 1c).

These data indicate that upregulation of BARD1protein is sufficient for p53 upregulation and phosphory-lation, suggesting that the function of BARD1 in p53stabilization involves phosphorylation of p53 at the keyresidue serine-15 and that increased BARD1 proteinlevels can induce this post-translational modification.To investigate whether BARD1 and p53 or phospho-

p53 physically interact, we used HEK 293T cells trans-fected with a cDNA expression vector of a BARD1–EGFP fusion construct expressed from the CMVpromoter (Jefford et al., 2004). Immunofluorescencemicroscopy was performed to visualize BARD1–EGFP,p53, and phospho-p53ser15 with specific antibodies. Whenantibodies recognizing total p53 were used, we observedthat all cells expressed p53 independently of BARD1–EGFP expression, whereas some accumulation of p53 innuclear dots was correlated with BARD1–EGFPexpression (Figure 2a). Using p53 antibodies specificfor phospho-serine-15, no staining was observed inuntransfected HEK 293T cells, while phospho-p53 wasexpressed in BARD1–EGFP-positive cells, andBARD1–EGFP and phospho-p53ser15 co-localized innuclear dots (Figure 2a). Comparison of p53 andphospho-p53ser15 in nontransfected cells and in

Figure 2 BARD1 co-localizes with phospho-p53. (a) HEK293T cells nontransfected or transfected with BARD1–EGFP expressionvector were used for immunofluorescence microscopy. DAPI staining, BARD1–EGFP fluorescence, anti-p53, and anti-phospho-p53staining are shown. Arrowheads mark cells with high BARD1–EGFP content, arrows mark those with moderated BARD1–EGFPexpression. Merge shows colocalization (arrow) of BARD1–EGFP and phospho-p53. (b) The histogram demonstrates BARD1, p53,and phospho-p53 expression in nontransduced and BARD1-transduced cells, derived from three independent transductionexperiments

BARD1 induces apoptosisA Feki et al

3

Oncogene

Page 98: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

BARD1–EGFP-expressing cells clearly shows thatphospho-p53 is only present in BARD1–EGFP expres-sing cells, and the amount of phospho-p53 labeling iscorrelated with the level of BARD1–EGFP expression.The quantitative analysis of BARD1–EGFP and phos-pho-p53ser15 expression shows similar percentage ofBARD1- and phospho-p53ser15-positive cells (Figure 2b).These results are consistent with the observation thatoverexpression of BARD1 can induce p53 phosphoryla-tion (Figure 1c).We have previously shown that BARD1 binds to p53

in co-immunoprecipitation assays (Irminger-Fingeret al., 2001). To test BARD1 binding to phospho-p53,immunoprecipitation experiments were performed withcell extracts from HEK 293T, TAC-2, and NuTu-19

cells, using a C-terminal BARD1 antibody (Figure 3a).To obtain increased protein levels of BARD1 andp53, HEK 293T and NuTu-19 cells were transducedwith BARD1-expressing lentiviral vector, and TAC-2cells were treated with doxorubicin. To quantify thetransduction efficiency, control transductions wereperformed using a GFP-expressing lentiviral vector. Asshown previously (Irminger-Finger et al., 2001), theincrease of BARD1 was associated with the increase ofp53 expression as compared to control transduction.Moreover, p53 and phospho p53ser-15 levels were elevatedin doxorubicin-treated TAC-2 cells and in BARD1-transduced HEK 293T and NuTu-19 cells. BARD1antibodies co-immunoprecipitated p53 and phosphop53ser-15 in all cell types. Phospho-p53ser-15 was co-

Figure 3 BARD1 interacts with p53 and with phospho-p53. (a) Co-immunoprecipitations were performed with HEK 293T, NuTu-19,and TAC-2 cells, using anti-BARD1 antibody C-20. TAC-2 cells, with or without doxorubicin treatment, and HEK293T and NuTu-19, nontransduced or transduced with BARD1 expression vector, were compared. Western blots were probed with anti-p53 antibodiesor anti-phospho epitope serine-15 antibodies. The left panel shows control immunoprecipitation supernatants, the right panel showspellets. (b) BARD1 stabilizes wild-type (wt) and triple mutant 33-81-315 (tri), but not p53 mutants of ser15 or ser37. Prostate cancercells PC3, deficient of p53 (Honda et al., 2002), were transfected (þ ) or not transfected (�) with BARD1 and cotransfected with p53 orp53 mutants. Western blots of cell extracts were probed with anti-p53 antibodies. (c) Region sufficient for p53 binding includes ANKand the region between ANK and BRCT. EGFP-tagged BARD1 and deletion constructs, and EGFP alone, were transientlytransfected. Immunoprecipitation was performed with anti-p53 antibodies and Western blots were probed with anti-GFP antibodies.The supernatant (S) and pellet (P) fractions show the fractions of the respective fusion proteins. (d) Schematic diagram of BARD1protein sequence and deletion-bearing fusion constructs is presented. Percentage of p53 binding, determined by gel intensity scan, andcapacity of apoptosis induction reported previously, (Jefford et al., 2004), are indicated. Arrows above the protein scheme indicaterelative positions of tumor-associated mutations (Thai et al., 1998; Ghimenti et al., 2002; Karppinen et al., 2004)

BARD1 induces apoptosisA Feki et al

4

Oncogene

Page 99: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

immunoprecipitated with BARD1 as efficiently as p53,albeit the amounts of p53 and phospho-p53ser-15 werereduced in NuTu-19 as compared to TAC-2 and HEK293T cells. These results indicate that BARD1 interactswith p53 and with phospho-p53ser-15 in different celltypes and cells from different species.To further investigate the specificity of BARD1-

dependent p53ser15 phosphorylation, we tested themutated forms of p53. Cotransfection assays were per-formed in p53-deficient human prostate cancer cells PC3(Honda et al., 2002). No p53 can be detected in PC3 cellswithout exogenous p53 expression, also ser15 and ser37mutants do not express detectable amounts of proteinafter transfection. Triple mutant 33-81-315, however, isstable (Zacchi et al., 2002). BARD1 cotransfection leadsto an increase of wild-type p53 and of the triple mutant,but had no effect on p53 mutated at ser15 and ser37(Figure 3b).The interaction of BARD1 and p53 was also tested

with BARD1–EGFP fusion constructs, using antibodiesagainst p53 to co-immunoprecipitate BARD1–EGFP,which was assayed with anti-EGFP antibodies onWestern blots (Figure 3c). To determine the specificregion of BARD1 required for this interaction, BARD1-EGFP deletion constructs were used in transfectionassays followed by immunprecipitations with anti-p53antibodies. Interestingly, all deletion constructs co-immunoprecipitated with anti-p53 except RING-EGFP,which represents the amino-terminal 89 amino acids ofBARD1 (Figure 3c, d). The minimal region required forbinding to p53 was deletion construct ANK-EGFP,which was previously identified as the minimal regionsufficient for apoptosis induction (Jefford et al., 2004).This finding suggests that the molecular pathway ofBARD1 signaling towards apoptosis is based on its rolein p53 stabilization by phosphorylation.Since p53 phosphorylation on serine 15 can be

induced by overexpression of BARD1, it was temptingto speculate that BARD1 might act as a mediatorbetween p53 and its specific kinase. Phosphorylation ofp53ser-15 and p53ser-37 is required for p53 stabilization andfor p53-dependent response to genotoxic stress (Shiehet al., 1997; Tibbetts et al., 1999; Jack et al., 2004).A role of BARD1–BRCA1 complexes in phosphoryla-tion of p53ser-15 in an ATM/ATR pathway has beendescribed recently (Fabbro et al., 2004). We suspectedthat DNA-PK kinase might be responsible for p53ser-15

phosphorylation in a BARD1-dependent pathway, sinceDNA-PK can phosphorylate p53ser-15 upon genotoxicstress (Woo et al., 1998, 2002) and DNA-PK andKu-70, the active subunit of DNA-PK, are upregulatedin ATM/ATR-dependent stress response (Brown et al.,2000; Shangary et al., 2000). Most recently, it was foundthat Ku-70 mutation partially suppresses the homo-logous-repair defects of BARD1 disruption (Stark et al.,2004), suggesting that both act in a common pathway.We performed co-immunoprecipitation assays with

anti-p53 antibodies, anti-Ku-70, the active subunit ofDNA-PK, or anti-BARD1 antibodies to investigate theinteractions between DNA-PK, p53, and BARD1(Figure 4a). Immunoprecipitation assays, when per-

formed on lysates from HEK 293T cells, or HEK 293Tcells transduced with BARD1–EGFP, revealed thatBARD1 and p53 were co-immunoprecipitated with Ku-70. When anti-p53 antibodies were used to precipitatep53 or interacting proteins, p53 immunoprecipitationwas nearly 100%, as was BARD1 co-immunoprecipita-tion with anti-Ku-70 antibodies (Figure 4a).We were interested whether NuTu-19 cells expressed

Ku-70 and immunoprecipitation experiments wereperformed with untransfected and BARD1-transfectedNuTu-19 cells (Figure 4b). Only exogenous BARD1could accumulate detectable amounts of Ku-70 in im-munoprecipitation. No precipitation was observed whenBARD1 antibodies were omitted.As further evidence for the implication of DNA-PK, a

member of the phosphoinositide 3-kinase related kinase(PIKK) family, in phosphorylation of p53 and inter-action with BARD1, we treated cells with caffeine anddoxorubicin. Caffeine is a well-known inhibitor ofPIKK kinase activity and as a result can effectivelyinhibit the phosphorylation of p53ser-15. Doxorubicin onthe other hand induces genotoxic stress on cells,consequently resulting in phosphorylation of p53 onthe same residues inhibited by caffeine. After treatmentof HEK 293T cells with 50mM caffeine alone, weobserved a slight elevation in p53 levels as compared tountreated cells; however, there was no change in thephospho-p53 levels (Figure 4c). Doxorubicin treatmentinduces the expression of both p53 and BARD1, asdescribed previously (Irminger-Finger et al., 2001;Jefford et al., 2004). Since HEK 293T cells haveconstitutive p53 expression, no significant changes inp53 levels with or without genotoxic stress wereobserved. We did not observe substantial increase inphospho-p53 protein levels co-immunoprecipitatingwith BARD1 when treated with caffeine, but whentreated with doxorubicin. The combined treatment withcaffeine and doxorubicin resulted in decreased phospho-p53 levels in both the supernatant and pellet fractions ofthe BARD1 immunocomplexes (Figure 4c). This resultprovides further evidence that p53 is phosphorylated bya PIKK family member and that this reaction might becatalysed by the presence of BARD1, since BARD1immunocomplexes contain both phosphorylated andunphosphorylated species of p53. The same immuno-precipitation extracts were probed with anti-ATMantibodies and no binding of ATM to BARD1 couldbe observed (Figure 4c), thus supporting that DNA-PKrather than ATM is responsible for BARD1-dependentp53ser-15 phosphorylation.While p53 phosphorylation was induced in several cell

types by exogenous expression of BARD1, this effectwas most striking in malignant ovarian cancer cellsNuTu-19. NuTu-19 cells were generated by continuousin vitro culture of ovarian epithelial cells leading tospontaneous malignant transformation (Rose et al.,1996). These cells lack full-length BARD1 protein(Figure 1b, c). Cloning and sequencing of the BARD1mRNAs expressed in NuTu-19 cells revealed that noneof the cDNAs encoded a full-length BARD1 gene(Figure 5a, c, d). The cDNA corresponding to the 2.2 kb

BARD1 induces apoptosisA Feki et al

5

Oncogene

Page 100: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

wild-type BARD1 transcript contained point mutationsleading to a premature translation stop at position 1786between the ankyrin repeats (ANK) and the BRCTdomains (Figure 5a). A smaller 2 kb fragment was dueto deletion of exons 2 and 3. This mRNA species, hereindesignated BARD1g, was similar to the testis-specificmRNA BARD1b (Feki et al., 2004). This finding wascorroborated by Western blots performed with threedifferent previously characterized anti-BARD1 antibo-dies (Feki et al., 2004), which showed that full-lengthBARD1 protein is absent in NuTu-19 cells (Figure 5b).HEK 293T expressed a 94-kDa protein, detectable withN-terminal antibodies H-300 and PVC, directed againstthe region adjacent to the RING finger, and JH3,directed against the regions between ANK and BRCT.None of the antibodies tested recognized a proteinfragment corresponding to the molecular weight of full-length BARD1 in NuTu-19 cell extracts. Consistently,the most abundant transcript of BARD1 in NuTu-19cells was an 800 bp RT–PCR product, a novel splicevariant, hitherto named BARD1d. BARD1d lacks exons2–7, which include regions coding for the RING finger

and the ankyrin repeats (Figure 5c, d). Therefore,NuTu-19 cells do not express full-length BARD1molecules and none of the BARD1 transcripts encodeANK repeats and the region between ANK and BRCT.Interestingly, nucleotides 1618–1641 represent an

inverted repeat of region 804–826, which could act asantisense to abrogate the translation of BARD1, but notBARD1d transcripts. Mutations and deletions in theBARD1 coding region were in strong contrast to thestatus of p53. Cloning and sequencing of the p53 cDNAfrom NuTu-19 cells revealed two polymorphisms, S66Pand L192F, within the less conserved domains of p53,but no mutations that predict structural changes of theencoded protein.We had evaluated the apoptotic activity of NuTu-19

cells after treatment with UV or doxorubicin. NuTu-19cells show little response to apoptosis-inducing drugsas measured by FACS and TUNEL assays. Doxorubi-cin treatment resulted in no significant increase ofapoptotic cells as observed after TUNEL or 7AADstaining and flow cytometry analysis (Figure 6a). How-ever, when we overexpressed wild-type BARD1 by

Figure 4 BARD1 interaction with p53 and DNA-PK subunit Ku-70. (a) HEK293T (293T) nontransduced or transduced withBARD1 were used for immunoprecipitation assays with Ku-70, p53, and BARD1 antibodies. Supernatants (S) and pellet (P) fractionsare shown. (b) NuTu-19 cells nontransduced or transduced with BARD1 were used to demonstrate BARD1–Ku-70 interaction. Littleco-precipitation is observed in the absence of exogenous BARD1. Control experiment without antibodies shows no precipitation(beads). (c) P53 phosphorylation is induced (doxo) or inhibited (caffeine). Immunoprecipitation performed with BARD1 antibody C-20 to show interaction between BARD1, phospho-p53, p53, and ATM. Supernatants (S) and pellet (P) fractions are shown

BARD1 induces apoptosisA Feki et al

6

Oncogene

Page 101: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

transient transfection, apoptosis was induced to asimilar extent in NuTu-19 cells as in nonmalignantcontrol cells, TAC-2. We also transduced NuTu-19 cells

with wild-type BARD1 cloned into a lentiviral vectorunder the transcriptional control of a tetracycline-inducible promoter. Induction of BARD1 by increasing

Figure 5 NuTu-19 cells lack full-length BARD1. (a) RT–PCR of p53 and BARD1 transcripts from NuTu-19 and control RT–PCRfrom TAC-2 cells. (b) Western blot of HEK 293T and NuTu-19 cells probed with anti-BARD1 antibodies H-300, PVC, and JH-3.Expression of full-length 97 kDa BARD1 is observed in HEK 293T but not in NuTu-19 cells. (c) Schematic presentation of BARD1protein with conserved regions RING, ankyrin (ANK) and BRCT, and epitopes for antibodies H-300, PVC, and JH-3 (l signs),aligned with RT–PCR products BARD1g and BARD1d. Triangles above indicate position of introns. Small arrows on BARD1g andBARD1d indicate position of inverted repeats. (d) Comparison of the deduced protein sequence of rat BARD1 with isoformsBARD1a, BARD1g, and BARD1d. Positions of point mutations leading to translation stop in BARD1a are indicated by asterisks,polymorphisms are in black. Boxed sequences indicate RING finger, ANK, and BRCT, respectively

BARD1 induces apoptosisA Feki et al

7

Oncogene

Page 102: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

doxycyclin levels clearly resulted in apoptosis in a dose-dependent manner (Figure 6b). However, BARD1expression and apoptosis in response to doxorubicin innontransduced NuTu-19 cells was comparable to back-ground levels of apoptosis induction in BARD1-transduced cells without induction of gene expressionby doxycyclin.Similarly, we tested p53-phosphorylation on p53ser15

as dose response to BARD1 expression levels in NuTu-19 cells. Nontransduced cells and cells transduced withincreasing amounts of BARD1 expression vectors werecompared on Western blots. Phospho-p53ser15 increasewas observed in a dose-dependent manner on Westernblots probed with antibodies against phospho-p53ser15

(Figure 6c).This experiment proved that expression of wild-type

BARD1 could reverse apoptosis resistance in cellslacking full-length BARD1. In nonmalignant cells with

endogenous wild-type BARD1, overexpression of exo-genous BARD1 is sufficient to stabilize p53 and toinduce p53ser-15 phosphorylation. In BARD1-deficientNuTu-19 cells, phosphorylation of p53 cannot beinduced by apoptosis-inducing drugs, but by theexogenous expression of wild BARD1, without otherstress signals, suggesting that BARD1 facilitates thephosphorylation of p53 by binding to p53 and Ku-70,the subunit of DNA-PK, and that this is sufficient andrequired for apoptosis induction.

Discussion

In the present study we demonstrate that BARD1 playsa critical role in phosphorylation of p53, which isrequired for its apoptotic function. Several kinases

Figure 6 Expression of BARD1 reverses apoptosis resistance in NuTu-19 cells. (a) TUNEL assay on NuTu-19 ovarian cancer cells ornonmalignant TAC-2 cells with or without treatment with doxorubicin (doxo) or transfection of BARD1. Percentage of TUNEL-positive cells was quantified by flow cytometry. (b) NuTu-19 cells and NuTu-19 transduced with lentiviral vector containing BARD1were monitored for apoptosis. Cells were incubated with or without doxycyclin to induce BARD1 expression, or with doxorubicin(doxo). (c) Dose response of BARD1-induced phosphorylation is observed after transduction of NuTu-19 cells with BARD1.Nontransduced (0), transduced (1� ), or transduced with threefold (3� ) concentration of BARD1 expression viral vector, cells wereprobed on Western blot with anti-phospho-ser15 p53-specific antibody

BARD1 induces apoptosisA Feki et al

8

Oncogene

Page 103: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

involved in p53 activation have been reported (Lu et al.,1997; Shieh et al., 1997, 1999; Tibbetts et al., 1999;Hirao et al., 2000) and ATM/ATR-dependent phos-phorylation of p53ser-15, implicating BRCA1–BARD1complexes, has been described very recently (Fabbroet al., 2004).It was shown before that BRCT domains, a protein

motif also present in BARD1 and BRCA1, havephospho-epitope-binding capacities (Rodriguez et al.,2003). A phospho-epitope-binding function was re-ported specifically for BRCT domains of BRCA1.BRCT domains are present in many other repairproteins and seem to facilitate physical interactionsamong proteins involved in the cellular response toDNA damage check point control and DNA repair(Bork et al., 1997; Callebaut and Mornon, 1997). It isspeculated that BRCT domains are important modulesfor tumor suppressor functions, supported by cancer-associated mutations within the BRCT domain ofBRCA1 (Williams et al., 2001).BARD1 contains tandem BRCT motifs, it has tumor

suppressor functions (Irminger-Finger and Leung,2002), and it binds and stabilizes p53 (Irminger-Fingeret al., 2001) which is suggestive of a role in a pathway ofp53 phosphorylation. Here we show that, indeed,BARD1 binds to p53 and phospho-p53 and, secondly,BARD1 interacts with the DNA-PK subunit, Ku-70,but not ATM. The region sufficient for p53 interactionand apoptosis induction comprises ANK, the regionbetween ANK and BRCT, and part of the BRCTdomains. It is possible that the BRCT domains arebinding to the kinase and that its target is binding to theadjacent region.Ku-70 is abundantly expressed in HEK 293T cells, as

is p53, as compared to BARD1. Nevertheless, over-expression of BARD1, without an additional stimulusthrough DNA damage, causes p53 phosphorylation inHEK 293T cells and NuTu-19 cells, suggesting thatBARD1 is an essential and limiting partner for theformation of the p53–DNA–PK complex. Consistentwith this view, BARD1 co-localizes with phospho-p53 indistinct nuclear dots (Figure 5b). Since overexpressionof BARD1 is sufficient for p53ser15 phosphorylation,binding of BARD1 to the kinase and its target catalysesp53 phosphorylation.Supporting this view, several cancer-associated mis-

sense mutations (Thai et al., 1998; Ghimenti et al., 2002;Ishitobi et al., 2003), map to the region defined sufficientfor p53 binding and apoptosis induction, and it isdeleted or mutated in the NuTu-19 cells, described here.It is likely that BARD1 deficiencies are frequent in

cancer cells that have wild-type p53 but are resistant toapoptosis, and significant changes in BARD1 expressionhave been found associated with malignancies (Pomeroyet al., 2002; Iyengar et al., 2003; Schafer et al., 2003;Zuco et al., 2003). The NuTu-19 ovarian cancer cells,to some extent, are a model of ovarian cancer, whichsuggests that BARD1 is a key target for mutationsleading to carcinogenesis. NuTu-19 cells are derivedfrom a spontaneous transformation of continuouslycultured ovarian epithelial cells (Rose et al., 1996),

suggesting that few genetic modifications led tothe phenotype of resistance to undergo apoptosis.This process mimics the in vivo situation, as it hasbeen hypothesized that proliferation associated withovarian repair contributes to the risk of ovarian cancerin humans (Fathalla, 1972). We speculated and confirmedhere that one of the critical mutations for malignanttransformation involves the tumor suppressor BARD1.Indeed, rat ovarian cancer cells, NuTu-19, have

applied several strategies to eliminate the expression offunctional BARD1: mutations leading to a truncatedform of the protein, differential splicing leading to thedeletion of essential parts of the protein, and possiblyrepression of the partially active forms of the protein byexpression of a short inverse repeat sequence that couldfunction as antisense RNA in transcriptional repressionof wild-type forms of BARD1. These mechanisms ofinhibition of BARD1 function are consistent with thehypothesis that BARD1 is an essential tumor suppres-sor, critical for apoptosis induction upon damage, bystabilizing p53 (Irminger-Finger et al., 2001). BARD1binding to p53 could also influence p53 stability byinhibiting its ubiquitination and/or its binding toMDM2 (Meek, 1994; Honda et al., 1997; Milczareket al., 1997).Consistently, no phosphorylation of p53 is observed in

NuTu-19 cells in the absence of exogenous expression ofwild-type BARD1. Co-immunoprecipitation experimentssuggest that BARD1 binds both the kinase and itsrespective target p53 and catalyses its phosphory-lation. This could be explained mechanistically by ahigher affinity between BARD1 and Ku-70 and betweenBARD1 and p53 than between p53 and Ku-70. Our datasupport this view, since co-immunoprecipitations ofBARD1 and p53 or BARD1 and Ku-70 are moreefficient than co-immunoprecipitations of p53 and Ku-70. In addition, BARD1 protein levels are considerablylower in the cell than either p53 or Ku-70, thereforeBARD1 can be regarded as the critical and limitingfactor in the formation of the p53/Ku-70 complexes.Furthermore, the overexpression of BARD1 alone with-out induction of the DNA damage pathway is sufficientfor p53 phosphorylation and supports the view thatBARD1 catalyses p53 phosphorylation by DNA-PK.

Materials and methods

Cell cultures

The HEK 293T cell line, as well as HeLa cells, were obtainedfrom the ATCC. NuTu-19 cells (Rose et al., 1996) were agenerous gift from Dr Attila Major, Geneva, PC3, humanprostate cancer cells, were obtained from Dr TimothyMacDonnell, Houston. All cell lines were cultured in 10 cmplastic tissue culture dishes (NUNC, Roskilde, Denmark) andincubated at 371C in humidified air containing 5% CO2.Adherent subcultures were detached from culture dishes usinga solution of Trypsin/EDTA in HBSS without Ca2þ or Mg2þ .Subconfluent cell cultures were passaged regularly at 1 : 10dilution approximately twice a week. All cell culture media,solutions, buffers, and antibiotics were purchased from

BARD1 induces apoptosisA Feki et al

9

Oncogene

Page 104: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

GIBCO (Invitrogen AG, Basel, Switzerland). HEK 293T cellswere pretreated for 1 h with 50mM caffeine, after which theywere either treated with doxorubicin (50 mg/ml) or leftuntreated as described previously (Irminger-Finger et al.,1998).

Western blotting

For Western blots, protein extracts from cell cultures wereprepared in standard lysis buffer (150mM NaCl, 0.1% SDS,50mM Tris (pH 8.0)). Equal concentrations of all whole cellextracts per gel were loaded onto SDS–PAA gels according toLaemmli’s protocol and were transferred to polyvinylidenefluoride (PVDF) membranes. Filters were blocked for 1 h inPBS containing 5% non-fat milk and then incubated at roomtemperature for 2 h with a polyclonal BARD1 (Irminger-Finger et al., 1998), p53 and P-p53 (Santa Cruz) antibodies.After washing, membranes were incubated with horseradishperoxidase-conjugated secondary antibody and the resultingcomplexes were visualized by an enhanced chemiluminescence(ECL) system (Boehringer).

Immunoprecipitation

Immunoprecipitation was performed by homogenizing cellsamples in 100 ml (100–150 mg total protein) RIPA buffer(150mM NaCl, 1% NP-40, 0.5% NaDOC, 0.1 SDS, 50mMTris (pH 8.0), and 2mM EDTA (pH 8.0)). Cells extracts wereincubated with p53 antibody (FL-393 Santa Cruz), or GFPantibody (Torrey Pines Biolabs, Inc) in a 1 : 200 dilution overnight, followed by incubation with sepharose-bound protein Gfor 3 h. Extracts were centrifuged, and pellets were washedtwice with RIPA buffer and re-suspended in 100ml RIPAbuffer. In all, 25ml of supernatant and pellet fractions wasanalysed by Western blotting. Total protein content insupernatant and pellet fractions was compared by fast greenstaining after protein transfer and was 90–10% at least.Antibody reactive bands were visualized, scanned, andquantified using Imagequant.

Antibodies

The anti-BARD1 (H300), anti-p53 (FL-393), and anti-phospho p53 ser15 antibodies (Cell Signaling 5284) werepurchased from Santa Cruz. Anti-GFP (TP-401) was pur-chased from Chemokine (Torrey Pines Biolabs, Houston, TX,USA). WFS is a previously described anti-BARD1 antibody(Gautier et al., 2000). Horseradish peroxidase (HRP)-con-jugated secondary antibodies were purchased from Santa Cruzand were compatible with Cruz Molecular Weight Markers(Santa Cruz, CA, USA).

RT–PCR

Total RNA was purified from NuTu-19 cells using Qiagenreagent. RNA was reverse transcribed using oligo-dT andsuperscript II (Life Technologies). RT products were subjected

to PCR analysis using specific primers for rat BARD1 cDNA(covering BARD1 coding regions) and rat p53 (covering p53coding regions). PCR cycles were: 941C for 15min, 941C for1min, 561C for 1min, 721C for 2min 30 s for BARD1 and1min 30 s for p53 (35� ), then 721C for 10min. Equal volumesof PCR products were analysed on 1% agarose gel.

Lentiviral vector production and transduction

The packaging constructs used in this study were thepCMVDR8.92 and the pCMVDR8.93 plasmids describedpreviously (Dull et al., 1998). The vesicular stomatitis virusG (VSV-G) envelop was the pM2DG plasmid (Zufferey et al.,1997). The BARD1 coding sequence was cloned into thecarrier pLOXEW plasmid (Zufferey et al., 1997). The viralparticles were produced using the usual transient transfectionmethod (Zufferey et al., 1997). The 293T cells (2.8� 106 cells in10-cm tissue culture dishes) were co-transfected with either10 mg of pLOXEW-BARD1 ires-GFP, 3.75mg ofpCMVDR8.92, 3.75 mg of pCMVDR8.93, and 2.5mg ofpM2DG, or 10mg of pHR0-CMV-BARD1, 3.75mg ofpCMVDR8.92, 3.75 mg of pCMVDR8.93, and 2.5mg ofpM2DG.After an overnight incubation in the presence of the

precipitate, the culture medium (10ml) was changed. Twodays later, the supernatant was harvested, filtered through0.45-mm pore-sized polyethersulfone membrane and subse-quently 1ml of each was added to NuTu-19 target cells in six-well plates (104 cells/well). Vector particles were left on the cellsfor 3 days. Transduction efficiency was monitored using aGFP-expressing vector.

Plasmid constructions and transfection

Different BARD1 mutants fused with GFP, was describedpreviously, were used. Briefly, the full-length mouse BARD1cDNA (Irminger-Finger et al., 1998) was used to generate thedeletion constructs in pcDNA3 behind a CMV promoter. Theplasmid pEGFP-N1 (enhanced green fluorescent protein) waspurchased from Clontech and was used to generate either thefull-length BARD1–EGFP or BARD1 deletion-bearing mu-tants fused in-frame with the EGFP tag. The deletion-bearingconstructs were the same as described (Jefford et al., 2004).HEK293T cells were seeded in 10 cm petri dishes andtransfected with 2 mg of DNA using the Effectene reagentsfrom Qiagen. Cells were harvested 48 h post-transfection.

AcknowledgementsWe are grateful to MH Sow, A Caillon, and C Genet fortechnical help. We are specifically indebted to Dr A Major forsupply and sharing of expertise in the NuTu-19 cell line. Weare grateful to G Del Sal for his generous gift of expressionplasmids of the p53 mutants and T McDonnell for PC3 cellline. This work was supported by grant 3100-068222 from theSwiss National Science Foundation to IIF and KHK.

References

Baer R and Ludwig T. (2002). Curr. Opin. Genet. Dev., 12,86–91.

Bork P, Hofmann K, Bucher P, Neuwald AF, Altschul SF andKoonin EV. (1997). FASEB J., 11, 68–76.

Brown KD, Lataxes TA, Shangary S, Mannino JL, GiardinaJF, Chen J and Baskaran R. (2000). J. Biol. Chem.., 275,6651–6656.

Callebaut I and Mornon JP. (1997). FEBS Lett., 400, 25–30.

Chehab NH, Malikzay A, Stavridi ES and Halazonetis TD.(1999). Proc. Natl. Acad. Sci. USA, 96, 13777–13782.

Dull T, Zufferey R, Kelly M, Mandel RJ, Nguyen M, Trono Dand Naldini L. (1998). J. Virol., 72, 8463–8471.

Fabbro M, Kienan S, Hobson K, Deans AJ, Powell SN,McArthur GA and Khanna KK. (2004). J. Biol. Chem., 279,31251–31258.

Fathalla MF. (1972). Obstet Gynecol. Surv., 27, 751–768.

BARD1 induces apoptosisA Feki et al

10

Oncogene

Page 105: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Feki A, Jefford CE, Durand P, Harb J, Lukas H, Krause K-Hand Irminger-Finger I. (2004). Biol. Reprod., 71, 1614–1624.

Gautier F, Irminger-Finger I, Gregoire M, Meflah K and HarbJ. (2000). Cancer Res., 60, 6895–6900.

Ghimenti C, Sensi E, Presciuttini S, Brunetti IM, Conte P,Bevilacqua G and Caligo MA. (2002). Genes ChromosomesCancer, 33, 235–242.

Hao M, Lowy AM, Kapoor M, Deffie A, Liu G and LozanoG. (1996). J. Biol. Chem., 271, 29380–29385.

Hashizume R, Fukuda M, Maeda I, Nishikawa H, Oyake D,Yabuki Y, Ogata H and Ohta T. (2001). J. Biol. Chem., 276,14537–14540.

Hirao A, Kong YY, Matsuoka S, Wakeham A, Ruland J,Yoshida H, Liu D, Elledge SJ and Mak TW. (2000). Science,287, 1824–1827.

Honda R, Tanaka H and Yasuda H. (1997). FEBS Lett., 420,25–27.

Honda T, Kagawa S, Spurgers KB, Gjertsen BT, Roth JA,Fang B, Lowe SL, Norris JS, Meyn RE and McDonnell TJ.(2002). Cancer Biol. Ther., 1, 163–167.

Iyengar P, Combs TP, Shah SJ, Gouon-Evans V, Pollard JW,Albanese C, Flanagan L, Tenniswood MP, Guha C,Lisanti MP, Pestell RG and Scherer PE. (2003). Oncogene,22, 6408–6423.

Irminger-Finger I and Leung WC. (2002). Int. J. Biochem. CellBiol., 34, 582–587.

Irminger-Finger I, Leung WC, Li J, Dubois-Dauphin M,Harb J, Feki A, Jefford CE, Soriano JV, Jaconi M,Montesano R and Krause KH. (2001). Mol. Cell, 8,

1255–1266.Irminger-Finger I, Soriano JV, Vaudan G, Montesano R andSappino AP. (1998). J. Cell. Biol., 143, 1329–1339.

Ishitobi M, Miyoshi Y, Hasegawa S, Egawa C, Tamaki Y,Monden M and Noguchi S. (2003). Cancer Lett., 200, 1–7.

Jack MT, Woo RA, Motoyama N, Takai H and Lee PW.(2004). J. Biol. Chem., 279, 15269–15273.

Jefford CE, Feki A, Harb J, Krause KH and Irminger-FingerI. (2004). Oncogene, 23, 3509–3520.

Jin Y, Xu XL, Yang MC, Wei F, Ayi TC, Bowcock AMand Baer R. (1997). Proc. Natl. Acad. Sci. USA, 94,

12075–12080.Joukov V, Chen J, Fox EA, Green JB and Livingston DM.(2001). Proc. Natl. Acad. Sci. USA, 98, 12078–12083.

Karppinen SM, Heikkinen K, Rapakko K and Winqvist R.(2004). J. Med. Genet, 41, e114.

Knippschild U, Milne DM, Campbell LE, DeMaggio AJ,Christenson E, Hoekstra MF and Meek DW. (1997).Oncogene, 15, 1727–1736.

Kohn KW. (1999). Mol. Biol. Cell, 10, 2703–2734.Lakin ND and Jackson SP. (1999). Oncogene, 18, 7644–7655.Levine AJ. (1997). Cell, 88, 323–331.Lohrum M and Scheidtmann KH. (1996). Oncogene, 13,

2527–2539.Lu H, Fisher RP, Bailey P and Levine AJ. (1997). Mol. Cell.

Biol., 17, 5923–5934.Ludwig T, Fisher P, Ganesan S and Efstratiadis A. (2001).

Genes Dev., 15, 1188–1193.Manke IA, Lowery DM, Nguyen A and Yaffe MB. (2003).

Science, 302, 636–639.

McCarthy EE, Celebi JT, Baer R and Ludwig T. (2003). Mol.Cell. Biol., 23, 5056–5063.

Meek DW. (1994). Semin. Cancer Biol., 5, 203–210.Milczarek GJ, Martinez J and Bowden GT. (1997). Life Sci.,60, 1–11.

Oda K, Arakawa H, Tanaka T, Matsuda K, Tanikawa C,Mori T, Nishimori H, Tamai K, Tokino T, Nakamura Yand Taya Y. (2000). Cell, 102, 849–862.

Pomeroy SL, Tamayo P, Gaasenbeek M, Sturla LM, AngeloM, McLaughlin ME, Kim JY, Goumnerova LC, Black PM,Lau C, Allen JC, Zagzag D, Olson JM, Curran T, WetmoreC, Biegel JA, Poggio T, Mukherjee S, Rifkin R, Califano A,Stolovitzky G, Louis DN, Mesirov JP, Lander ES andGolub TR. (2002). Nature, 415, 436–442.

Rodriguez M, Yu X, Chen J and Songyang Z. (2003). J. Biol.Chem., 278, 52914–52918.

Rose GS, Tocco LM, Granger GA, DiSaia PJ, Hamilton TC,Santin AD and Hiserodt JC. (1996). Am. J. Obstet. Gynecol.,175, 593–599.

Schafer R, Tchernitsa OI, Zuber J and Sers C. (2003). Adv.Enzyme Regul., 43, 379–391.

Scully R, Chen J, Ochs RL, Keegan K, Hoekstra M, FeunteunJ and Livingston DM. (1997). Cell, 90, 425–435.

Shangary S, Brown KD, Adamson AW, Edmonson S, Ng B,Pandita TK, Yalowich J, Taccioli GE and Baskaran R.(2000). J. Biol. Chem., 275, 30163–30168.

Shieh SY, Ikeda M, Taya Y and Prives C. (1997). Cell, 91,325–334.

Shieh SY, Taya Y and Prives C. (1999). EMBO J., 18,

1815–1823.Stark JM, Pierce AJ, Oh J, Pastink A and Jasin M. (2004).

Mol. Cell. Biol., 24 (21), 9305–9316.Thai TH, Du F, Tsan JT, Jin Y, Phung A, Spillman MA,Massa HF, Muller CY, Ashfaq R, Mathis JM, Miller DS,Trask BJ, Baer R and Bowcock AM. (1998). Hum. Mol.Genet., 7, 195–202.

Tibbetts RS, Brumbaugh KM, Williams JM, Sarkaria JN,Cliby WA, Shieh SY, Taya Y, Prives C and Abraham RT.(1999). Genes Dev., 13, 152–157.

Williams RS, Green R and Glover JN. (2001). Nat. Struct.Biol., 8, 838–842.

Woo RA, Jack MT, Xu Y, Burma S, Chen DJ and Lee PW.(2002). EMBO J., 21, 3000–3008.

Woo RA, McLure KG, Lees-Miller SP, Rancourt DE and LeePW. (1998). Nature, 394, 700–704.

Wu LC, Wang ZW, Tsan JT, Spillman MA, Phung A, Xu XL,Yang MC, Hwang LY, Bowcock AM and Baer R. (1996).Nat. Genet., 14, 430–440.

Yu X, Chini CC, He M, Mer G and Chen J. (2003). Science,302, 639–642.

Zacchi P, Gostissa M, Uchida T, Salvagno C, Avolio F,Volinia S, Ronai Z, Blandino G, Schneider C and Del Sal G.(2002). Nature, 419, 853–857.

Zuco V, Supino R, De Cesare M, Carenini N, Perego P, GattiL, Pratesi G, Pisano C, Martinelli R, Bucci F, Zanier R,Carminati P and Zunino F. (2003). Biochem. Pharmacol., 65,1281–1294.

Zufferey R, Nagy D, Mandel RJ, Naldini L and Trono D.(1997). Nat. Biotechnol., 15, 871–875.

BARD1 induces apoptosisA Feki et al

11

Oncogene

Page 106: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

97

Results submitted for publication

5. BARD1 interacts with BRCA2 to regulate cytokinesis and maintain genomic stability Jefford CE, Delaval B, Ryser S, Birnbaum D, Irminger-Finger I.

BRCA1, BRCA2 and BARD1 are caretaker proteins and bona fide tumour

suppressors since mutation in either of these genes is sufficient for tumour initiation.

Furthermore, a knockout of either of the aforementioned gene in mice induces early

embryonic death due to proliferation defects and not to apoptosis. Tumour cells with

mutated BRCA1, BRCA2 or BARD1 exhibit definite chromosomal instability;

repression of either of these genes in cultured cells by knockdown experiments

causes genomic instability, the hallmark of all tumour cells. In this study we

investigate further the mechanisms that link BARD1-mediated tumour suppression

and maintenance of genomic stability. It was shown by us and others that BARD1

expression is not only augmented upon genotoxic stress, but is also regulated during

cell cycle. Here, we show by antibody staining that BARD1 is highly expressed

during mitosis in several cell lines and it co-localizes with microtubule spindle. We

showed that this staining is specific since the polyclonal antibody used recognizes

the BARD1-EGFP constructs with the N-terminal domain and not the RING deleted

BARD1-EGFP recombinants. Knockdown of BARD1 expression by transducing a

anti-BARD1 shRNA results in slower cell cycles, delayed cytokinesis, and

incompletion of mitosis. Furthermore, this phenotype is often accompanied by failure

of efficient chromosome segregation as indicated by the appearance of aberrant

karyotypes. Finally, we showed that BARD1 co-immunoprecipitates with Aurora B,

but not with Aurora A. For the first time we reveal an interaction of BARD1 with

BRCA2 in the regulation of cytokinesis, since BARD1 co-immunoprecipitates with

BRCA2 during the G2/M phase of the cell cycle. These data give us new evidence

that BARD1 may regulate cell-cycle progression through the Aurora B kinase

pathway for completion of cytokinesis. The mechanisms involved in cell-cycle

regulation and cytokinesis are not fully understood, but could be controlled by the

ubiquitin ligase activity of BARD1/BRCA1 dependant on AuroraB phosphorylation.

Cytogenetic analysis was provided by Sarantis Gagos, while some of the IPs were

done in collaboration with the Birenbaum laboratory.

Page 107: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Jefford et al. 1

Nature Cell Biology Letter

Inhibition of cytokinesis after BARD1 depletion causes genomic instability

Charles Edward Jefford1, Bénédicte Delaval2, Stephan Ryser1#, Sarantis Gagos3#, Daniel Birnbaum2, and Irmgard Irminger-Finger1*

1Biology of Aging Laboratory, Department of Geriatrics, Geneva University Hospitals, Geneva, Switzerland 2Institut de Cancérologie de Marseille, Laboratoire d'Oncologie Moléculaire, UMR599 Inserm-Institut Paoli-Calmettes, Marseille, France 3Laboratory of Genetics, Foundation of Biomedical Research, Academy of Athens, Greece # These authors were contributing equally *to whom correspondence should be addressed: Dr. Irmgard Irminger-Finger Biology of Aging Laboratory Dept. of Rehabilitation and Geriatrics University and University Hospitals Geneva Chemin de Petit Bel Air 2 CH-1225 Geneva/Chêne-Bourg Switzerland Tel: +41 22 305 54 57 Fax: +41 22 305 54 55 E-Mail: [email protected] http://www.medecine.unige.ch/recherche/biologie_du_vieillissement/

Key words: BARD1, BRCA1, BRCA2, TACC1, Aurora B, cytokinesis, chromosomal instability

Page 108: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Jefford et al. 2

Mutations in BARD1 (BRCA1 associated Ring protein1) are associated with breast and ovarian

cancer2-5, and BARD1 repression leads to genomic instability6-8. Genomic instability is also the

hallmark of cancers with in BRCA1 and BRCA2 mutations9. The BRCA1-BARD1 complex

controls centrosome amplification through ubiquitin ligase activity10, and BRCA2 is involved in

regulation of cytokinesis11, both mechanisms that are required for accurate chromosome

segregation. Here we report that BARD1, overexpressed during mitosis12,13, is localized to the

mitotic spindle and to the midbody at telophase and cytokinesis. In mitotic extracts BARD1

interacts with TACC1, Aurora B, and BRCA2, proteins with similar location and repression

phenotypes at cytokinesis11,14. siRNA depletion of BARD1 leads to Aurora B stabilization on the

midbody and inhibition of abscission of inter-sister cell microtubule bridges, associated with

numerical and structural chromosomal instability. These findings suggest that BARD1 is

required for regulation of Aurora B displacement/stability in a complex of BARD1-AuroraB-

BRCA2 that controls progression through cytokinesis.

A common feature of BRCA1, BRCA2, and BARD1 knock-out mice is embryonic lethality8, 15-

18, underscoring that these genes are vital for a variety of cellular events10, 19. During mitosis,

BRCA2 localizes to the midbody where it is involved in cytokinesis, as demonstrated by siRNA

depletion of BRCA2, which impedes cell separation11. In contrast to BRCA2, BRCA1 is

localized to centrosomes at anaphase onset, where in association with BARD1, it acts as E3

ubiquitin-ligase on γ-tubulin10. This is substantiated by the fact that abrogation of BRCA1 leads

to excessive centrosome duplication, incurring genomic instability10, 20. BARD1 is the major

protein binding partner of BRCA11 , and some of the functions of BARD1 are tightly linked to

the ubiquitin ligase activity of the BARD1/BRCA1 heterodimer21. However, independently of

BRCA1, BARD1 upregulation mediates apoptosis induction by catalyzing p53

phosphorylation22, 23.

Page 109: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Jefford et al. 3

Upregulated expression of BARD1 is also observed in mitotic cells12, 13, suggestive of its

implication in mitotic mechanisms. To further characterize BARD1 expression and function in

mitosis, indirect immunofluorescence was performed on non-synchronized cell populations.

Consistently BARD1 levels were increased in mitotic cells (Fig. 1A). BARD1 staining was

accumulated around and along the mitotic spindle (Fig. 1B). This intracellular localization of

BARD1 during mitosis was observed when using different anti-BARD1 antibodies in different

cell lines and across species, such as in human HeLa, MCF-7 (data not shown), PC3, and mouse

non-malignant mammary gland TAC-2 cells6. Omission of the primary antibody or competition

with blocking peptides did not lead to significant staining. The specificity of BARD1 staining

was further confirmed with anti-BARD1 antibody H-300, directed against the N-terminal 300

amino acids, on cells expressing exogenous BARD1-EGFP fusion proteins or BARD1-EGFP

deletion constructs (Support data Fig. 1A). Co-localization of anti-BARD1 staining and EGFP-

fluorescence was observed with full length BARD1 and ∆-RING-EGFP but not with ∆-HIND-

EGFP, a deletion of the N-terminal coding half, or EGFP alone. This demonstrates that BARD1

antibodies specifically recognize epitopes in the amino terminus of BARD1 (Support data Fig.

1B).

To further ascertain the localization of BARD1 on mitotic microtubules (MT), we performed

double-immuofluorescence staining with antibodies against BARD1 and against β-tubulin in

three different cell lines. Clearly, BARD1 colocalizes with spindle MT, as demonstrated in

metaphase cells (Fig. 1C). The localization to spindle MT is nearly exclusive in HeLa cells, but

additional cytoplasmic staining is observed in PC3 and to some extent in TAC-2 cells. BARD1

concentration on MT seems much higher than the tubulin concentration; or, anti-BARD1 staining

might be competitive for anti-tubulin staining, since the tubulin stain is less intense when the

anti-BARD1 stain is strong in double labeling experiments. The upregulation of BARD1 during

mitosis might be controlled at the protein level by CDK2, which regulates its ubiquitin ligase

function13.

Page 110: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Jefford et al. 4

We then investigated the association of BARD1 with MT at different stages of the cell cycle.

BARD1 localizes to spindle MT at all stages of cell cycle (Fig. 2). While nuclear dots and some

cytoplasmic staining could be observed in typical interphase cells (Fig. 2A, S-phase merge), most

BARD1 was localized to MT emanating from the spindle poles at metaphase and anaphase.

BARD1 and β-tubulin co-localization was observed as yellow staining, and BARD1 and DNA

co-localization as pink staining (Fig. 2B), indicating that a fraction of BARD1 co-localized with

MT and a small fraction with chromosomes. BARD1 and MT co-localization was specifically

observed at anaphase (Fig. 2B). Excessive BARD1 staining was also observed at the spindle

midzone and at cytokinesis at the midbody (Fig. 2A-B). While BARD1 staining was more

abundant than tubulin staining from anaphase to telophase, similar intensities were observed at

cytokinesis. During the furrowing of the cleavage plane, and until the end of cytokinesis, BARD1

staining intensified along the spindle midzone and even more so at the midbody.

These localizations of BARD1 at different stages of mitosis suggest that BARD1 migrates along

MT from the spindle pole to the midzone and accumulates at the midbody. While at onset of

anaphase, BARD1 could act with BRCA1 as ubiquitin ligase on γ-tubulin10, functions at later

stages of mitosis could be suspected.

Among the proteins that also localize to the midzone during anaphase and to the midbody during

cytokinesis, are BRCA211 and transforming acidic coiled-coil-containing protein (TACC1)14.

TACC proteins are involved in MT stabilization and abnormal centrosome formation, and have

been implicated in cancer progression (reviewed in24). TACC1 colocalizes and forms a functional

complex with the mitotic kinase Aurora B during cytokinesis14. Aurora B depletion by siRNA-

mediated interference prevents the formation of the midbody, consequently affects TACC1

localization to the midbody, and leads to abnormal cell division and multinucleated cells, a

phenotype resembling BARD1 antisense depletion6. Since BARD1 localizes to the spindle

midzone at anaphase and at the midbody during telophase and cytokinesis, we asked whether

Page 111: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Jefford et al. 5

BARD1 interacted with TACC1 and Aurora B. TACC1 as a MT stabilizing protein could have

an adaptor function for localization of BARD1 on MT.

Co-immunoprecipitation experiments were performed with anti-TACC1 and anti-BARD1

antibodies. When probing HeLa cell extracts, from non-synchronized cells or cells blocked at

different stages of the cell cycle, BARD1 co-precipitated with TACC1 in cells blocked at the

G1/S or at the G2/M border (Fig. 3A). When co-immunoprecipitation experiments were

performed with TACC1 antibodies and Western blots were probed with anti-BARD antibodies,

BARD1 was co-immunoprecipitated with TACC1 mostly in mitotic cells (Fig. 3B). An increase

of BARD1 was observed in mitotic cell extracts. BARD1 protein appeared as a double band of

97 kD and slightly larger, which could represent the phosphorylated and unphosphorylated

forms13. The fraction of BARD1 co-precipitating with TACC1 in mitotic cells corresponded to

the lower, presumably unphosphorylated form. No co-immunoprecipitation of BARD1 was

found in the control assay with beads only. The reverse experiment showed that anti-BARD1

antibodies co-precipitated TACC1 (Fig. 3B). In contrast to BARD1, TACC1 expression was not

increased during mitosis and only a fraction of TACC1 was co-immunoprecipitated with

BARD1, but the interaction was restricted to cells arrested in G2/M. As for BARD1, TACC1

antibodies recognized an additional band with slightly higher molecular weight, which was

increased in G2/M total lysates (Fig. 3A-B*). These results show that BARD1 and TACC1

interact in G2/M phase of the cell cycle and suggest a role for both proteins in a common

pathway.

It was shown previously that TACC1 is coexpressed temporally and spatially with Aurora B, a

mitotic kinase, critical for progression through mitosis25, 26. We therefore tested the BARD1-

Aurora B interaction. Western blots of anti-BARD1 immunoprecipitations were probed with

antibodies against Aurora A and B (Fig. 3C). Both Aurora A and B levels were increased in G1/S

and more in G2/M blocked cell extracts. However, co-immunoprecipitation with anti-BARD1

Page 112: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Jefford et al. 6

antibodies was only observed in G2/M cell extracts and only for Aurora B. This observation is

consistent with the specific subcellular localization of TACC1 and Aurora B and their potential

roles at the exit of mitosis14, in contrast to the localization of Aurora A, which acts with

BRCA127 at the spindle poles in centrosome duplication. The BRCA2 protein had also been

localized to the midbody11 and we therefore tested coimmuneprecipitation of BRCA2 with

BARD1. BRCA2 was binding to BARD1, in mitotic extracts, while the bona fide binding partner

BRCA1 was binding to BARD1 in cell extracts of cells blocked at G1/S and at G2/M. This data

suggest that during mitosis BARD1 sequentially interacts with BRCA1 and BRCA2. The specific

interaction of BARD1 and BRCA2 in mitotic extracts suggests that BARD1 acts in a pathway

involving TACC1, Aurora B, and BRCA2.

To assess the role of BARD1 during mitosis, we abrogated BARD1 expression in HeLa, MCF-7,

PC-3, and TAC-2 cells with siRNA and monitored the resulting phenotype. Immunofluorescence

microscopy was performed on cells transciently transfected with BARD1 siRNA expressed from

a constitutive promoter, using anti-BARD1 and anti–tubulin antibodies. When harvesting cells 36

hours after transfection, we observed a striking phenotype of cells cycle arrest in mitosis,

specifically an accumulation of cells at late telophase (Fig. 4A). MT staining showed that

separated sister cells at cytokinesis were connected by MT bridges, while exhibiting decondensed

chromosomes. The number of cells connected by midbody bridges was ten fold increased in the

siRNA treated cells, as compared to non-treated cells (Fig. 4B). Depletion of BARD1 was

specific, since siRNA against mouse BARD1 had not effect, compared to mock or human

siRNAs, as shown on Western blots (Fig. 4C). Residual expression of BARD1 is not found on

MT bridges when double labeling with anti-tubulin and anti-BARD1 antibodies was performed

(Fig. 4D) as compared to wild type cells (Fig. 2B). Thus, the depletion of BARD1 by siRNA

leads to a delay or arrest of cell cycle in cytokinesis and suggests that BARD1 might be required

for proper abscission of inter-sister cells MT bridges.

Page 113: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Jefford et al. 7

We investigated the localization of TACC1, Aurora B and BRCA2 in BARD1 depleted cells and

found striking differences in expression and localization for Aurora B (Fig. 4E), but not TACC1

or BRCA2 (data not shown). Aurora B localizes to distinct dots on inter-sister cells MT bridges,

bracketing the abscission point14. Unlike the specific and exclusive localization of Aurora B to

MT bridges at cytokinesis in wild type cells, in BARD1 repressed cells Aurora B dots are larger

and their location is less precise. Aurora B sustained on the cell membrane of daughter cells after

cytokinesis (Fig. 4E), while in wild type cells Aurora B dislocates to interphase nuclei of sister

cells. No differences were observed for Aurora B staining at other stages than cytokinesis. These

findings suggest that BARD1 controls Aurora B stability at the exit of mitosis.

Since the BARD1 depletion phenotype is reminiscent of the Aurora B depletion phenotype,

which also causes aneuploidy, it is suggestive that the interaction of BARD1 with Aurora B is

functionally significant. Malfunction of chromosome segregation due to a delayed and abnormal

cell division, could be caused by a deficient Aurora B- BARD1 complex.

The phenotype of BARD1 deficiency is also similar to the phenotype induced by siRNA

inhibition of BRCA2, in particular the phenotype of the BRCA2 mutation Capan-111. Like the

Capan-1 mutation, BARD1 siRNA inhibition leads to the accumulation of cells with elongated

midbodies. BARD1 and BRCA2 localize to the same regions during the late stages of mitosis

and it is likely that interaction of BARD1 with Aurora B and BRCA2 are required for completion

of cytokinesis.

The function of BARD1 in abscission of inter-sister cells MT bridges provides an explanation for

the observed genomic instability in BARD1-deficient cells6-8. Inhibition or delay of cytokinesis

could cause genomic instability due to improper chromosome segregation. To test this hypothesis

we performed a profound cytogenetic analysis of BARD1 expressing and BARD1 depleted HeLa

and MCF-7 cells.

Page 114: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Jefford et al. 8

Using alpha-satellite probes specific for chromosomes 7, 17, 18, and X, and locus specific probes

for chromosomal positions 13q12 and 21q12, we found that BARD1 repression leads to

increased rates of chromosomal losses and overall numerical deviation in MCF-7 or HeLa cells

(Table 1). We also observed masked numerical chromosomal instability due to whole

chromosome losses compensated by non-disjunctions in MCF-7 cells using dual color FISH

specific for centromeres 7 and 17 (Fig. 5). The rates of numerical instability of centromeres 7 and

17 were twice as pronounced in the BARD1 depleted MCF-7 cells than in BARD1 expressing

cells. Taken together these results show that loss of BARD1 leads to numerical chromosomal

instability, presumably due to the defective abscission of inter-sister cells MT bridges, which is

reminiscent of defects in mitotic fidelity as a consequences of Aurora-B ablation14.

The similarity of genomic instability phenotypes induced by the depletion of BARD1, BRCA1,

or BRCA2 is only partially reflected in the specific functions that were attributed of each of these

proteins. However, the sequential interaction of BARD1, first with BRCA1 at the centrosome as

ubiquitin ligase on γ-tubulin, a mechanism of control of centrosome duplication, and secondly

with BRCA2 at cytokinesis, provides an explanation for their overlapping functions resulting in

similar phenotypes. As negative regulator of centrosome amplification, an obvious cause of

aneuploidy28, 29, BARD1 acts as tumor suppressor with BRCA1. In regulation of cytokinesis,

BARD1 and BRCA2, together with Aurora B and TACC1, are required to ascertain proper cell

division and chromosome segregation to avoid aneuploidy. Hence the dual role of BARD1 in

mitosis, by interacting with BRCA1 and BRCA2, could explain the common phenotype of

genomic instability induced by BARD1, BRCA1 or BRCA2 deficiencies.

Page 115: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Jefford et al. 9

Materials and Methods

Cell Culture

TAC-2 cells are a clonal sub-population derived from the NMuMG normal murine mammary

gland epithelial cell line (CRL 1636; American Type Culture Collection, Rockville, MD) and

were used previously 6. They were cultured in collagen-coated tissue culture dishes (Falcon,

Becton Dickinson, San Jose, CA) in high glucose DMEM (GIBCO, Invitrogen corporation)

supplemented with 10% fetal calf serum (FCS, GIBCO), penicillin (110 µg/ml) and streptomycin

(110 µg/ml). The Human Embryonic Kidney HEK 293T cell line, as well as HeLa cell lines,

were obtained from the American Tissue Type Culture Collection. The PC3 cell line was

obtained from McDonnell. All cell lines were cultured in the aforementioned tissue culture

medium (TCM) containing low glucose concentration, using 9 cm plastic tissue culture dishes

and incubated at 37°C in humidified air containing 5% CO2. Adherent subcultures were detached

from culture dishes using a solution of Trypsin/EDTA (Life Technologies, Inc.). Sub-confluent

cell cultures were passaged regularly at 1:10 dilution twice a week.

Transfections of siRNA constructs

Cells were seeded in either 6 well plates or in 10 cm petri dishes and transfected with either

0.4µg or 2µg of DNA using the Effectene reagents provided by Qiagen. Cells were harvested at

24 and 48 hours post-transfection. They were centrifuged and washed once in PBS, resuspended

in 0.3 ml of ice cold radio immuno-precipitation assay (RIPA) buffer [150mM NaCl, 1% Nonidet

P-40, 0.5% NaDOC, 0.1% SDS, 50mM Tris, pH 8.0, 2mM EDTA, pH 8.0] and supplemented

with protease inhibitors [1mM PMSF, 3µg/ml aprotinin, 10µg/ml leupeptin] and phosphatase

inhibitors [2mM NaF, 25 mM activated Na3VO4]. Cells were lysed on ice for 30 minutes and

vortexed at regular intervals. The extracts were centrifuged at 13'000 rpm in a microcentrifuge

for 15 minutes at 4°C in order to remove cell debris and clear the cell lysate. The human BARD1

Page 116: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Jefford et al. 10

siRNA construct was generated by annealing the following complementary oligonucleotides and

inserted into the pSuper vector as a BglII/HindIII fragment: forward, 5'-G ATC CCC CAT TCT

GAG AGA GCC TGT GTT CAA GAG ACA CAG GCT CTC TCA GAA TGT TTT TGG

AAA-3', and reverse, 5'-AG CTT TTC CAA AAA CAT TCT GAG AGA GCC TGT GTC TCT

TGA ACA CAG GCT CTC TCA GAA TGG GG-3'. This construct was then transiently

transfected in Hela cell as described. Mouse siRNAs were directed against the respective

homologous regions.

Antibodies

The anti-BARD1 (H-300), anti-Aurora A and B, antis-BRCA1 and 2 antibodies were purchased

from Santa Cruz. Anti-GFP (TP-401) was purchased from Chemokine (Torrey Pines Biolabs).

PVC and WSF were described previously6. Anti-TACC1 antibody was purchased from Upstate

Biotechnology Euromedex (Mundolsheim, France).

Secondary conjugated antibodies to horseradish peroxidase (HRP) were purchased from Santa

Cruz and were compatible with Cruz Molecular Weight Markers (Santa Cruz). Secondary

immunoflurescent donkey anti-mouse, rabbit and goat conjugated to FITC or Alexa red 555

fluorochromes were obtained from Molecular Probes.

Immunofluorescence microscopy

Cells were seeded on glass cover slips in 6 well culture dishes or plated on a 8 well coverslide

and transfected or not with the aforementioned plasmids. Cells were washed in HBSS, and either

fixed 2% paraformaldehyde (PFA) in Hank's Balanced Salt Solution (HBSS) 15 min at room

temperature (R.T.) or in methanol for 6 minutes at -20c and rinsed in acetone for 30 seconds.

PFA-fixed cells were permeabilized in 1% Triton/PBS for 15 minutes at R.T. and thereafter

blocked for in 1% serum/PBS for 30 min. Coverslides were incubated with appropriate

Page 117: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Jefford et al. 11

antibodies for 1 hour at R.T. in 1%FCS/PBS, washed and stained with DAPIC for 3 min. Finally

coverslides were mounted with fluorogard solution, observed under a Nikon epi-fluorescence

microscope and images captured with a 3.3 mega-pixel CCD camera. Images were processed

with Metamorph software (Visitron).

Immunoprecipitation

Protein lysates used for immunoprecipitation correspond to HeLa cells synchronized (G1/S),

synchronized and released (G2/M) and not synchronized (NS) using a double thymidine block.

24 h after plating, cells were arrested in S-phase using 2mM thymidine for 18 h, released from

the arrest for 9 h, arrested a second time with thymidine, and released again to obtain a

population enriched in mitotic cells. Proteins from lysates were precipitated with anti-TACC1 or

anti-BARD1 antibodies, and bound proteins were detected with anti-BARD1, anti-TACC1, anti-

Aurora A or B, or anti-BRCA1 or 2 antibodies.

Interphase dual color cytogenetic analysis

To investigate numerical instability we applied dual-color interphase FISH, which is more

accurate than conventional karyotype analysis, satellite probes specific for chromosomes 7, 17,

18, and X were used. We also applied locus specific probes hybridizing to the Rb region of

chromosome 13 and the region critical for Down syndrome on chromosome 21. Probes were

either purchased from Vysis, or produced and tested in house, using an indirect labeling Nick

Translation process (ROCHE) according to the manufacturer’s protocol. Cell cultures of high

mitotic index were exposed to colcemid (0.1 µg/ml) (Gibco, BRL) for 1,5 h, at 37ºC and

harvested according to routine cytogenetic protocols. FISH was based on pepsin pre-treatment,

formamide target denaturation, over-night hybridization, and stringency post hybridization

Page 118: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Jefford et al. 12

washes. For each chromosome pair studied (i.e. chromosome a and chromosome b), a modal

number (Ma, Mb) was calculated as the most representative number of spots per nucleus per 100

cells/harvest. Chromosome loss was defined as Mn-1, chromosome gain as Mn+1. Overall

deviation was calculated as the number of nuclei that do not show modal number of spots.

Metaphase dual colour cytogenetic analysis Masked numerical chromosomal instability due to coupled chromosomal losses and non-

disjunctions was explored in the MCF-7 cells using the same dual-color technique as described

above. For this cytogenetic approach we performed conventional and multicolor FISH Karyotype

analysis to select a specific pair of chromosomes that showed at least two different recombinant

chromosomes of unequal size that could be easily identified on chromosome spreads. For MCF-7

cells centromeres 7 and 17 were a suitable pairs (Fig. 5). We deduced the stability of

recombinant and intact chromosomes, by FISH and inverted DAPI, in 50 randomly selected

compact metaphase spreads derived from BARD1 expressing or BARD1 depleted cells.

Chromosomal losses compensated by non-disjunction of one of the remaining homologous

centromeres, were recorded as two events of numerical instability/metaphase.

Page 119: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Jefford et al. 13

Figure legends

Figure 1. BARD1 expression during mitosis. (A) BAR1 is upregulated during in mitotic cells

(m). (B) BARD1 immunostaining of three different cell lines, TAC-2, HeLa, and PC-3. Indirect

immunofluorescence images show that BARD1 accumulates around mitotic spindles during

mitosis in all cell lines examined. BARD1 protein was flagged with H-300 antibody. Nuclei were

stained with DAPI. (C) BARD1 co-localizes with beta-tubulin staining in various cell lines. (A)

Mouse mammary gland TAC-2 and (B) prostate cancer PC3 cells were stained with anti-β-

tubulin labeled and anti-BARD1 (H-300) antibodies. Tubulin is labeled with FITC, BARD1 with

Alexa red 555, and nuclei are stained with DAPI.

Figure 2. Localization of BARD1 during stages of mitosis. (A) Immunostaining of an

asynchronous population of PC-3 cells with anti-BARD1 (H-300) and anti-β-tubulin showing

high immunoreactivity of BARD1 surrounding microtubule spindles at all stages of mitosis.

Tubulin is colored green and BARD1 in red. At anaphase BARD1 is mostly localized to the

spindle pole, only partial colocalization with tubulin. (B) At metaphase colocalization of BARD1

and tubulin staining marked in yellow (short arrow) and abundant BARD1 staining at the spindle

periphery (long arrow) can be observed. At telophase some of the colocalization with tubulin,

marked yellow (short arrow) is observed, but abundant staining of BARD1 is localized to the

midbody (long arrow), colocalization of BARD1 and DNA can be observed in pink (arrowhead).

In cytokinesis the most prominent BARD1 staining is seen at the midbody, colocalized with

tubulin (yellow).

Figure 3. BARD1 interacts with TACC1 and Aurora B, but not Aurora A, and BRCA2 but not

BRCA1 in mitotic extracts. The co-immunoprecipitation assay shows that TACC1 can pull down

Page 120: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Jefford et al. 14

full length BARD1 (C). Cells were either not enriched (--), enriched in S phase (S) with a double

thymidine block or enriched in G2/M phase of the cell cycle (G2/M) released after double

thymidine block. Western blot of total cell extract clearly shows that BARD1 is enriched during

mitosis, as compared to the lower unrelated protein which serves as loading control. The reverse

experiment shows that BARD1 can pull-down TACC1 (B), Aurora B, but not Aurora A (C),

BRCA1 in G1/S and G2/M extracts and BRCA2 in G2/M extracts only (D).

Figure 4. HeLa cells transiently transfected with BARD1 siRNA expressing plasmid. Cells were

collected and stained 20 hours post transfection with anti-BARD1 (H-300) and anti-β-tubulin

antibodies. Panel (A) demonstrates phenotype obtained with BARD1 repression, and percentage

of cells, delayed or arrested at late cytokinesis, with abnormal abscission of midzone body is

shown in (B). The fraction of cells with abnormal abscission is about 10 fold higher in BARD1

depleted cells than in the control cells transfected with mock siRNA. Residual BARD1 staining

is not localized to the midbody (D). Aberrant localization of Aurora B in BARD1 depleted cells

(E). Arrow indicates persistent Aurora B staining in sister cells after mitosis.

Figure 5. Numerical chromosomal instability in BARD1 depleted cells. Centromere specific

staining with probes for chromosomes 7 (orange) and 17 (green) was applied on MCF-7 cells

depleted of BARD1 by siRNA expression. Chromosomal losses are monitored by centromere

staining of interphase nuclei (a). Embedded partial karyotypes, representing the modal (most

representative) imbalances of chromosomes 7 and 17, were defined as chromosomal markers m1,

m2, m3, and m4. m1 and m2 were clonaly retained recombinant chromosomes derivatives of

chromosome 7. m3 and m4 were clonaly retained recombinant chromosomes derivatives of

chromosome 17 (G-Banding/Inverted DAPI FISH; magnification x1000) (b). Staining of four

spots per centromere of chromosomes 7 and 17 is representative for majority of metaphase

spreads (c). Loss of m4 and a translocation involving m3 is shown in (d). A non-disjunction of

Page 121: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Jefford et al. 15

recombinant m3 leads to increased number of spots (five) for centromere 17 (e). Examples of

masked numerical instability are shown in f-h: a loss of m4 is compensated by non-disjunction of

recombinant m3 (f), non-disjunction of recombinant m1 coupled to loss of one of the two

apparently intact chromosomes 7 (g), and loss of m1, coupled to non-disjunction of m2 (h).

Arrows indicate chromosomal non-disjunctions.

Support Figure 1. (A) Antibody control on cells transfected with various constructs. (B)

Schematic view of constructs.

Acknowledgements

We are grateful to T. McDonnell for the gift of PC3 cell line. We highly appreciate interesting

suggestions and observations made by A. Feki, and discussions with I. Bondarev and with P.

Berardi. We are indebted to A. Caillon and M. Hiourea for excellent technical assistance. This

work was supported by Swiss National Science Foundation grant 3100-068222 to I. I.-F. , Inserm

and Cancéropôle grants to D. B.; Foundation of Biomedical Research of the Academy of Athens

Greece starting grant to S.G.. B. D. is supported by a fellowship from Ligue Nationale Contre le

Cancer.

Page 122: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Jefford et al. 16

HeLa

Chromosome

Loss% Chromosome Gain% Overall deviation%

BARD1 (+) (-) (+) (-) (+) (-) 13 3 13 3 7 8 25 18 12 35 29 15 52 63 21 26 38 2 7 31 48 X 15 8 6 17 23 35

MCF-7

Chromosome

Loss% Chromosome Gain% Overall deviation%

BARD1 (+) (-) (+) (-) (+) (-) 7 10 19 1 2 12 22 13 9 10 14 14 26 28 17 20 33 2 6 23 44 18 11 30 0 20 21 46 21 23 17 14 3 46 45 X 16 20 1 1 18 21

Table 1: Percentages of chromosomal instability of BARD1 expressing (+) vs BARD1 depleted

cells (-).

Page 123: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Jefford et al. 17

References

1. Wu, L.C. et al. Identification of a RING protein that can interact in vivo with the BRCA1

gene product. Nat Genet 14, 430-440 (1996). 2. Thai, T.H. et al. Mutations in the BRCA1-associated RING domain (BARD1) gene in

primary breast, ovarian and uterine cancers. Hum Mol Genet 7, 195-202 (1998). 3. Ghimenti, C. et al. Germline mutations of the BRCA1-associated ring domain (BARD1)

gene in breast and breast/ovarian families negative for BRCA1 and BRCA2 alterations. Genes Chromosomes Cancer 33, 235-242 (2002).

4. Ishitobi, M. et al. Mutational analysis of BARD1 in familial breast cancer patients in Japan. Cancer Lett 200, 1-7 (2003).

5. Karppinen, S.M., Heikkinen, K., Rapakko, K. & Winqvist, R. Mutation screening of the BARD1 gene: evidence for involvement of the Cys557Ser allele in hereditary susceptibility to breast cancer. J Med Genet 41, e114 (2004).

6. Irminger-Finger, I., Soriano, J.V., Vaudan, G., Montesano, R. & Sappino, A.P. In vitro repression of Brca1-associated RING domain gene, Bard1, induces phenotypic changes in mammary epithelial cells. J Cell Biol 143, 1329-1339 (1998).

7. Joukov, V., Chen, J., Fox, E.A., Green, J.B. & Livingston, D.M. Functional communication between endogenous BRCA1 and its partner, BARD1, during Xenopus laevis development. Proc Natl Acad Sci U S A 98, 12078-12083 (2001).

8. McCarthy, E.E., Celebi, J.T., Baer, R. & Ludwig, T. Loss of Bard1, the heterodimeric partner of the Brca1 tumor suppressor, results in early embryonic lethality and chromosomal instability. Mol Cell Biol 23, 5056-5063 (2003).

9. Grigorova, M., Staines, J.M., Ozdag, H., Caldas, C. & Edwards, P.A. Possible causes of chromosome instability: comparison of chromosomal abnormalities in cancer cell lines with mutations in BRCA1, BRCA2, CHK2 and BUB1. Cytogenet Genome Res 104, 333-340 (2004).

10. Starita, L.M. et al. BRCA1-dependent ubiquitination of gamma-tubulin regulates centrosome number. Mol Cell Biol 24, 8457-8466 (2004).

11. Daniels, M.J., Wang, Y., Lee, M. & Venkitaraman, A.R. Abnormal cytokinesis in cells deficient in the breast cancer susceptibility protein BRCA2. Science 306, 876-879 (2004).

12. Jefford, C.E., Feki, A., Harb, J., Krause, K.H. & Irminger-Finger, I. Nuclear-cytoplasmic translocation of BARD1 is linked to its apoptotic activity. Oncogene 23, 3509-3520 (2004).

13. Hayami, R. et al. Down-regulation of BRCA1-BARD1 ubiquitin ligase by CDK2. Cancer Res 65, 6-10 (2005).

14. Delaval, B. et al. Aurora B -TACC1 protein complex in cytokinesis. Oncogene 23, 4516-4522 (2004).

15. Gowen, L.C., Johnson, B.L., Latour, A.M., Sulik, K.K. & Koller, B.H. Brca1 deficiency results in early embryonic lethality characterized by neuroepithelial abnormalities. Nat Genet 12, 191-194 (1996).

16. Hakem, R. et al. The tumor suppressor gene Brca1 is required for embryonic cellular proliferation in the mouse. Cell 85, 1009-1023 (1996).

17. Liu, C.Y., Flesken-Nikitin, A., Li, S., Zeng, Y. & Lee, W.H. Inactivation of the mouse Brca1 gene leads to failure in the morphogenesis of the egg cylinder in early postimplantation development. Genes Dev 10, 1835-1843 (1996).

18. Ludwig, T., Chapman, D.L., Papaioannou, V.E. & Efstratiadis, A. Targeted mutations of breast cancer susceptibility gene homologs in mice: lethal phenotypes of Brca1, Brca2, Brca1/Brca2, Brca1/p53, and Brca2/p53 nullizygous embryos. Genes Dev 11, 1226-1241 (1997).

Page 124: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Jefford et al. 18

19. Jasin, M. Homologous repair of DNA damage and tumorigenesis: the BRCA connection. Oncogene 21, 8981-8993 (2002).

20. Hsu, L.C., Doan, T.P. & White, R.L. Identification of a gamma-tubulin-binding domain in BRCA1. Cancer Res 61, 7713-7718 (2001).

21. Baer, R. & Ludwig, T. The BRCA1/BARD1 heterodimer, a tumor suppressor complex with ubiquitin E3 ligase activity. Curr Opin Genet Dev 12, 86-91 (2002).

22. Irminger-Finger, I. et al. Identification of BARD1 as mediator between proapoptotic stress and p53-dependent apoptosis. Mol Cell 8, 1255-1266 (2001).

23. Feki, A. et al. BARD1 induces apoptosis by catalysing phosphorylation of p53 by DNA-damage response kinase. Oncogene (2005).

24. Gergely, F. et al. The TACC domain identifies a family of centrosomal proteins that can interact with microtubules. Proc Natl Acad Sci U S A 97, 14352-14357 (2000).

25. Ducat, D. & Zheng, Y. Aurora kinases in spindle assembly and chromosome segregation. Exp Cell Res 301, 60-67 (2004).

26. Adams, R.R., Maiato, H., Earnshaw, W.C. & Carmena, M. Essential roles of Drosophila inner centromere protein (INCENP) and aurora B in histone H3 phosphorylation, metaphase chromosome alignment, kinetochore disjunction, and chromosome segregation. J Cell Biol 153, 865-880 (2001).

27. Ouchi, M. et al. BRCA1 phosphorylation by Aurora-A in the regulation of G2 to M transition. J Biol Chem 279, 19643-19648 (2004).

28. Brodie, S.G. & Deng, C.X. BRCA1-associated tumorigenesis: what have we learned from knockout mice? Trends Genet 17, S18-22 (2001).

29. Xu, X. et al. Centrosome amplification and a defective G2-M cell cycle checkpoint induce genetic instability in BRCA1 exon 11 isoform-deficient cells. Mol Cell 3, 389-395 (1999).

Page 125: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

A

Figure 1A-B

B

HeLa TAC-2PC-3m

m

m

Page 126: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

BARD1β-Tubulin Merge

TAC-2

DNA

PC-3

HeLa

Figure 1C

C

Page 127: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Telophase CytokinesisMetaphaseS-phase Anaphasem

erge

BAR

D1

β-tu

bulin

DN

AA

Figure 2A

Page 128: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

BTelophase CytokinesisAnaphase

Figure 2B

Page 129: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

175

BARD1

Cell lysate

NS G1/S G2/M

IP anti-TACC1 Beads only

NS G1/S G2/M NS G1/S G2/M

*

A

TACC1

Cell lysateIP anti-BARD1 Beads only

NS G1/S G2/M NS G1/S G2/M NS G1/S G2/M

*

B

C

Figure 3

83

Aurora B

Aurora A

32,5

DBRCA2

BRCA1 180

Page 130: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

0.38

3

0

0.5

1

1.5

2

2.5

3

3.5

Mock siRNA BARD1

% a

bnor

mal

absc

issi

on

Microtubule BARD1 Nucleus

A B

D

Figure 4A-D

C siBARD1 - m h

α-BARD1

α-Actin

Page 131: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Figure 4E

E wildtype

siRNA BARD1

Page 132: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

Figure 5

Page 133: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

124

Supplemental results and ongoing projects

6. Role of BARD1 in telomere maintenance and crisis Introduction

The BARD1 and BRCA1 proteins are involved in several mechanisms that maintain

genomic stability, because mutation or absence in either of these genes leads to

chromosomal instability and pre-cancerous lesions. Most cancer cells are prone to

genomic instability which is often correlated with a reactivated telomerase, whereas

normal somatic cells bear a repressed telomerase. BARD1 and BRCA1 form nuclear

dots during S-phase that are relocated upon DNA damage induced by hydroxyurea.

Furthermore, it was shown that BARD1 and BRCA1 aggregate in PML bodies, in

telomerase negative cell lines, and that the BRCA1-Mre11-Rad50-Nbs1 complex,

which intervenes during DNA repair, is also present in telomeric complexes.

Telomeres are ribonucleoprotein complexes poised on the end of chromosomes that

maintain the stability of the chromosomes.

HeLa

MCF-7 PC3

TAC-2

a

e

dc

f g

b

HeLa

MCF-7 PC3

TAC-2

a

e

dc

f g

b

Figure 6.1. Co-localisation of telomeres and BARD1. a) CHO cells stained with anti-TRF-2. b) Confocal image of HeLa cells stained with anti-BARD1 forming the characteristic nuclear foci. c) FISH of telomeric probe by using PNA probe (DAKO), red dots represent telomeres. d) PNA hybridisation in BARD1-/- mouse embryos. e) The four panels represent double staining with PNA in green and anti-BARD1 in red in four different cell lines. These are overlay images, therefore co-localisation is in yellow and is indicated with yellow arrows. f) and g) 293T cells transfected with full-length BARD1-EGFP and stained with anti-TRF-2 in f) and with anti-BARD1 as a positive control in g). Nuclei are stained with DAPI in blue.

BARD1-EGFP

Alexa Red555

Merge

α-TRF-2 α-BARD1

Page 134: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

125

1% 0,25%

Input mock

pc-α-BARD1 Ab

H300N19 C20 JH3 α-T

RF2

Hela

TAC2

A

n.s.bl.0

2000

4000

6000

8000

10000

12000

14000

16000

18000

20000

input 1%

input 0,25

%mock

mock-pc

N19 C20 JH3

H300

TRF2

ChIP samples

sign

alte

lom

eric

prob

e [a

.u.]

HelaTAC2

B

1% 0,25%

Input mock

pc-α-BARD1 Ab

H300N19 C20 JH3 α-T

RF2

Hela

TAC2

A1% 0,25%

Input mock

pc-α-BARD1 Ab

H300N19 C20 JH3 α-T

RF2

1% 0,25%

Input mock

pc-α-BARD1 Ab

H300N19 C20 JH3 α-T

RF2

Hela

TAC2

A

n.s.bl.0

2000

4000

6000

8000

10000

12000

14000

16000

18000

20000

input 1%

input 0,25

%mock

mock-pc

N19 C20 JH3

H300

TRF2

ChIP samples

sign

alte

lom

eric

prob

e [a

.u.]

HelaTAC2

n.s.bl.0

2000

4000

6000

8000

10000

12000

14000

16000

18000

20000

input 1%

input 0,25

%mock

mock-pc

N19 C20 JH3

H300

TRF2

ChIP samples

sign

alte

lom

eric

prob

e [a

.u.]

HelaTAC2

B

The absence of telomeres or of proteins responsible for maintaining telomere

maintenance results in chromosomal instability and inaccurate segregation of

chromosomes that perpetrate breakage-fusion-bridge cycles. Telomeres and their

associated proteins such as TRF-1 and TRF-2 (Telomere Repeat binding Factor 1

and 2) form dot-like structures as seen by FISH (Fluorescent in Situ Hybridisation)

experiments and by immunofluorescence respectively. In view of all the above

evidence, we wanted to see if there was a link between BARD1, telomeres and

chromosome instability. We investigated whether BARD1 repression would have an

effect on telomere length or on telomerase activity and whether BARD1 can interact

within telomeric complexes either by binding to telomeric DNA or to telomere

proteins.

Results and discussion

Since BARD1 and telomeres form

nuclear dots and that they share

some of the same proteins in their

complexes, we first investigated

whether BARD1 and telomeres would

be part of the same biochemical

complex and would co-localize

together in double immunofluorescent

experiments. In order to observe this

co-localisation, we performed indirect

immunofluorescence on paraffin

sections from mouse embryos and in

cultured cells with anti-BARD1

antibody combined with FISH using a

telomere probe (PNA labelling system

from DAKO). It is possible to observe

some co-localization between BARD1

and the PNA signal in certain places

but it is not flagrant (Figure 6.1.e).

Figure 6.2. Telomere Chromatin Immunoprecipitation assays (Telo-ChIP). The several anti-BARD1 antibodies (N19, C20, JH3, H300) are incapable of immunoprecipitating telomeric DNA, as compared to TRF-2 suggesting that there is no interaction between BARD1 and telomeric chromatin in either HeLa or TAC2 cells.

Page 135: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

126

We also performed PNA telomere hybridisation on sections of BARD1-/- embryos in

order to evaluate the state of the telomere structures. It appears that BARD1

knockouts display a non-altered PNA stain compared to wild type mouse embryos

(Figure 6.1.d). To further look at an interaction between BARD1 and telomeric DNA,

we performed chromatin immunoprecipitation assays by using anti-BARD1 antibody

and labelled telomeric DNA. The results are unfortunately not conclusive, since there

is no strong signal indicating an attachment of BARD1 on the telomeric DNA (Figure

6.2). Furthermore, to see if BARD1 interacts with components of the telomere

Figure 6.3. Immunohistochemistry showing BARD1 expression. Paraffin embedded sections in the spleen and in the liver of wild type (G0) and 5th generation of mTR-/- knockout mice (G5) were immunostained with anti-BARD1 antibody. Left column represents sections from wild type tissue and show no reactivity. The two columns on the right side are from telomerase KO mice. Represented in brown the reactivity of N19 anti-BARD1 antibody secondary antibody is conjugated to HRP, DAB staining. Top two rows are taken from paraffin embedded liver sections and bottom row from spleen. Slides were counterstained with Hemalin.

5x

5x 20x

60x

5x

G0 wild type G5 mTR -/-

40x 5x

5x 20x

Page 136: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

127

complex, we performed double fluorescent immunohistochemistry using anti-TRF-2

or hTERT and anti-BARD1 antibodies on cultured cells. Here again, only a few spots

of co-localization can be observed, but these slides need to be visualized with a

confocal microscope in order to get a more precise image. In order to assess the

interaction between BARD1 and TRF-2 or hTERT, we immuno-stained cells that

were expressing a BARD1-EGFP fusion protein with TRF-2 antibody (Figure 6.1.f).

Telo-Flow-FISH

0

50

100

150

200

250

300

350

TAC-2 ABI

Hum

an C

E ly

mph

ocyt

es

U2O

S

SAO

S

DNA

ladd

erM

W

TAC

-2

Hum

an lo

w c

ont

Hum

an L

E ly

mph

ocyt

es

TAC

2/A

BI

Hum

an h

igh

cont

Telomere length Flow-FISH

020406080

100120140160180

U2OS U2OS-hTERT

JURKAT Fibroblasts SAR13

Cell lines

Rela

tive

Fluo

resc

ence

A

C

B

21kb

Telo-Flow-FISH

0

50

100

150

200

250

300

350

TAC-2 ABI

Hum

an C

E ly

mph

ocyt

es

U2O

S

SAO

S

DNA

ladd

erM

W

TAC

-2

Hum

an lo

w c

ont

Hum

an L

E ly

mph

ocyt

es

TAC

2/A

BI

Hum

an h

igh

cont

Telomere length Flow-FISH

020406080

100120140160180

U2OS U2OS-hTERT

JURKAT Fibroblasts SAR13

Cell lines

Rela

tive

Fluo

resc

ence

A

C

B

21kb

Figure 6.4. Measurement of telomere length in normal and BARD1 repressed cells. Measurements were made by telomere-Flow-FISH and by southern blot of terminal restriction fragments (TRF) probed with a DIG labelled telomere probe. A) TAC2 and BARD1 repressed TAC2/ABI do not show significant differences. Cells with FITC labelled telomeres by in situ hybridisation are analysed by flow cytometry. FITC signal per cell is proportional to telomere length. Usually 10’000 events are counted. Signal is normally subtracted from non-labelled cells representing background signal. B) TAC2 and TAC2/ABI cells as well as control cells were assayed for telomere length by TRF southern. This assay displays physical length of cells. It is more direct than Flow-FISH. However, TAC2 and TAC2/ABI have marginally the same telomere length. Mouse telomeres are double the size of human telomeres. Human telomere length averages between 4 and 15 kb, whereas mouse telomeres are about 20-40kb long. C) U2OS and SAOS2 are telomerase negative cell lines that have very long telomeres compared to Jurkat, fibroblasts or even other ovarian cancer cells (SAR13).

Page 137: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

128

We observed that TRF-2 anti-body co-localizes to some of the dots formed by the

BARD1-EGFP fusion protein. These results show partial co-localization of TRF-2 and

BARD1. In the future, we would like to use the BARD1-EGFP recombinant for

biochemical experiments and for co-immunoprecipitation studies by using the various

proteins involved in telomere metabolism. These proteins include TRF-1, Rap-1,

hTERT, and so on. We also examined the expression of BARD1 in tissues from

telomerase knockout mice (mTR-/-). It is worth noticing that BARD1 expression is

increased in mTR-/- tissues, possibly because mTR-/- mouse tissues are more prone

to apoptosis. These mice show premature aging and telomere attrition (Figure 6.3).

In order to ascertain the effect of the absence of BARD1 on telomere length, we

performed flow-FISH and TRF on BARD1 repressed cells (Figure 6.4.). Flow-FISH is

an indirect labelling method to measure telomere length by fluorescent in situ

hybridisation quantified by flow cytometry. TRF (terminal restriction fragment assay)

utilizes southern blot method with a highly digested genomic DNA using a telomere

probe to measure the physical length of telomeres. Since the BARD1 repressed cell

lines did not give a clear result, we transfected primary cells and transformed cell

lines with a siRNA of BARD1. We measured telomere length by using flow-FISH. Our

preliminary results suggest that BARD1 repression could have an effect on telomere

length, since telomere length appears to be diminished in siRNA transfected cells

(Figure 6.5). These cells were assayed for telomerase activity by TRAP (telomere

repeat amplification protocol); a very sensitive method which amplifies by PCR the

telomere repeats added by telomerase elongation. However, no changes were

observed on telomerase expression (Figure 6.6). Therefore, telomere attrition

observed after BARD1 knock-down is not due to lack of telomerase activity.

Furthermore, we transduced a short hairpin interfering BARD1 RNA (shRNA) by

using a lentiviral system to obtain a constitutive and an inducible down regulation of

BARD1. These cells have not been assayed yet for telomere length attrition. Several

passages need to be created to observe an effect over time. Overall, our results do

not allow to dress a complete picture of BARD1 interfering in telomere metabolism.

However, these results do not exclude an indirect action of BARD1 on telomeres.

Page 138: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

129

SAR

2x

nega

tive

cont

TAC

-2

MC

F-7

10bp

DN

A la

dder

MW

TAC

-21

+ si

RN

A

HeL

Apo

sitiv

e co

ntSA

R n

egat

ive

cont

MC

F-7

+siR

NA

10bp

DN

A la

dder

MW

H2O

CH

APS

Buf

fer

HeLAHeLA

+siRNA

Buf

fer

TAC

2 p1

6

ABI p

8

TAC

2 p3

9

ABI p

30

RM

C

SAR

2x

nega

tive

cont

TAC

-2

MC

F-7

10bp

DN

A la

dder

MW

TAC

-21

+ si

RN

A

HeL

Apo

sitiv

e co

ntSA

R n

egat

ive

cont

MC

F-7

+siR

NA

10bp

DN

A la

dder

MW

H2O

CH

APS

Buf

fer

HeLAHeLA

+siRNA

10bp

DN

A la

dder

MW

H2O

CH

APS

Buf

fer

HeLAHeLA

+siRNA

Buf

fer

TAC

2 p1

6

ABI p

8

TAC

2 p3

9

ABI p

30

RM

C

Buf

fer

TAC

2 p1

6

ABI p

8

TAC

2 p3

9

ABI p

30

RM

C

Telomere Flow-FISH

0

20

40

60

80

100

120

Jurkat NuTu-19 TAC-2 HeLA MCF-7

Cell Lines

Rel

ativ

e Fl

uore

scen

ce

(-)

(+)siRNA

TAC2

HeLa

MCF7

siRNA

M 1x 2x 3x C -

BARD1

GAPDH

A BTelomere Flow-FISH

0

20

40

60

80

100

120

Jurkat NuTu-19 TAC-2 HeLA MCF-7

Cell Lines

Rel

ativ

e Fl

uore

scen

ce

(-)

(+)siRNA

Telomere Flow-FISH

0

20

40

60

80

100

120

Jurkat NuTu-19 TAC-2 HeLA MCF-7

Cell Lines

Rel

ativ

e Fl

uore

scen

ce

(-)

(+)siRNA

TAC2

HeLa

MCF7

siRNA

M 1x 2x 3x C -

BARD1

GAPDH

A B

Figure 6.5. Telomere attrition in BARD1 knock-down cells. TAC2, HeLa and MCF-7 cells were transiently transfected with siRNA of BARD1 3 times over a course of 10 days. A) RT-PCR shows BARD1 mRNAlevels are decreased relatively efficiently. B) Shown in black the relative length of telomeres from the cell lines which were treated with siRNA BARD1.

Figure 6.6. Measurement of telomerase activity by TRAP assay (Telomere Repeat Amplification Protocol). This PCR based assay amplifies telomeric repeats in presence of active telomerase. It appears that TAC2 cells are positive for telomerase at low passage (left panel). HeLa and MCF-7 cells transfected with siRNA BARD1 do not have repressed telomerase activity (middle and left panel).

Page 139: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

130

7. Characterization of the BARD1 promoter It was established that BARD1 protein and mRNA are increased during genotoxic

stress and during cell death. BARD1 is also expressed under hormonal control.

Furthermore, it appears that BARD1 stabilizes BRCA1 and vice versa. This

regulation is interdependent. In order to investigate the mechanisms regulating

BARD1 expression under various physiological environments, such as hormonal

treatment, inflammation, genotoxic insult, embryogenesis or cell death, we decided to

isolate and characterize the BARD1 promoter region. The first step was to identify the

5’UTR and the site of the initiation of transcription (cap site). To this purpose, we

performed 5’RACE and a so called “steplike RT-PCR” with forward primers at various

positions upstream of the BARD1 ATG site. We amplified the cDNA of BARD1 with

primers positioned at –114 upstream of the ATG, but not at position –192,

demonstrating that BARD1 initiation site is situated within this region. A position

situated further upstream of what has been previously published (Figure 7.1).

ATG = + 1

CAG

Exon 1

Initiation start site

Exon 2 Exon 3 Exon 4CTGCTC

? ?

- 28

position forward primers

- 73

- 280 - 114- 192 + 246+ 139

reverse primers

- 153 - 132

F 1=90 R 1=72F 2=91F 3=92F 4=93 R 2=94

Reverse primer R 1 (+139)

Reverse primer R 2 (+ 246)

F1 F2 F3 F4 F1 F2 F3 F4 F1 F2 F3 F4 F1 F2 F3 F4

Gen DNA HeLa Lympho CTB

HeLa Lympho CTB

H2O

H2O

F1 F2 F3 F1 F2 F3 F1 F2 F3

RT-PCR

- 864F 5=41

ATG = + 1

CAG

Exon 1

Initiation start site

Exon 2 Exon 3 Exon 4CTGCTC

? ?

- 28

position forward primers

- 73

- 280 - 114- 192 + 246+ 139

reverse primers

- 153 - 132

F 1=90 R 1=72F 2=91F 3=92F 4=93 R 2=94

Reverse primer R 1 (+139)

Reverse primer R 2 (+ 246)

F1 F2 F3 F4 F1 F2 F3 F4 F1 F2 F3 F4 F1 F2 F3 F4

Gen DNA HeLa Lympho CTB

HeLa Lympho CTB

H2O

H2O

F1 F2 F3 F1 F2 F3 F1 F2 F3

RT-PCR

- 864F 5=41

Figure 7.1. Identification of 5’UTR of BARD1 in humans. A) This figure depicts the “steplike-RT-PCR” that amplifies, with upstream primers of ATG, the cDNA derived from the BARD1 messenger. B) and C) RT_PCR using the indicated primers showing a step formation. F3 and F4 do not amplify any product.

B

A

C

Page 140: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

131

We first isolated a primary region of 865 base pairs upstream. In a second step, we

isolated a larger region of 2402 base pairs upstream of the BARD1 ATG by PCR on

genomic DNA on two healthy young adults. We cloned these two regions into the

pGL3 basic vector containing the luciferase reporter gene (Figure 7.2). This region

was sequenced and analysed for candidate transcription factors (Figure 7.3). The

luciferase gene is placed under the transcriptional control of the supposed BARD1

promoter. The vector with the BARD1 promoter region was then assayed for

luciferase activity by transfecting the pGL3b-BARD1-promoter plasmid into several

human and mouse cell lines (Figure 7.4).

HindIII

HindIII

A

B

-865 bp

-2402 bp

HindIII

HindIII

A

B

-865 bp

-2402 bp

Figure 7.2. Cloning of putative BARD1 promoter region. Two fragments were cloned into pGL3 basic vector. BARD1 promoter of -865 and -2402 base pairs upstream from first ATG cloned in MCS of pGL3 basic. The promoter region of upstream of BARD1 was cloned from human lymphocytes from 2 different origins, male and female. They were inserted in-sense and anti-sense into pGL3 basic vector and sequenced.

Page 141: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

132

After transfection in various cell lines, the BARD1 region of 865 base pairs appears to

be sufficient for constitutive activity of BARD1 since there is a high signal in all cell

lines. However, this region does not seem to have a regulatory potential since when

the transfected cells were treated with doxorubicin to induce a genotoxic stress, no

significant increase of the reporter luciferase was noticed (Figure 7.5). Furthermore,

when transfected cells were treated with several hormones no change in signal was

observed (Figure 7.6). In order to eliminate biological variation due to cell death or

transfection discrepancies, these experiments were performed using a dual luciferase

reporter system. The secondary Renilla luciferase serves as an internal control (pRL-

TK). It is transfected at the same moment as the reporter vector containing promoter

and is constitutively expressed under a thymidin kinase (TK) promoter, whereas the

reporter gene is a Firefly luciferase. The results are then normalized according to the

internal expression for each sample. Furthermore, for accuracy, the transfections are

performed in triplicate and measured at three points each.

luciferase

luciferase

luciferase

luciferase

-8-865-2411 2402-CE(+)

856-LE5(+)

856-CE1(+)

856-LE5(-)

ATG =1

5’ 3’-200-800-1800 -1600 -1400 -1200 -1000 -600 -400-2000-2200-2400

Putative initiation site CAG = - 74

GC box -90- Sp1 and GC rich

HSF2HSF2

NF-κB

GATA-3 GATA-2

MZF1MZF1MZF1 MZF1

p300

C/EBPC/EBP

C/EBPC/EBP

C/EBPaC/EBPb

RORα1 AhR/Ar

USF USF USF USF

AML1-aAML-1aAML-1a

AML-1aAML-1a

Oct-1 Sp-1

SRY SRY SRY SRY AP-1 AP-1 AP-1

GATA-2

p53

luciferase

luciferase

luciferase

luciferase

-8-865-2411 2402-CE(+)

856-LE5(+)

856-CE1(+)

856-LE5(-)

luciferase

luciferase

luciferase

luciferase

-8-865-2411 2402-CE(+)

856-LE5(+)

856-CE1(+)

856-LE5(-)

ATG =1

5’ 3’-200-800-1800 -1600 -1400 -1200 -1000 -600 -400-2000-2200-2400

Putative initiation site CAG = - 74

ATG =1

5’ 3’-200-800-1800 -1600 -1400 -1200 -1000 -600 -400-2000-2200-2400

Putative initiation site CAG = - 74

GC box -90- Sp1 and GC rich

HSF2HSF2 HSF2HSF2

NF-κB

GATA-3 GATA-2

MZF1MZF1MZF1 MZF1

MZF1MZF1MZF1 MZF1

p300

C/EBPC/EBP

C/EBPC/EBP

C/EBPaC/EBPb

RORα1 AhR/Ar

USF USF USF USFUSF USF USF USF

AML1-aAML-1aAML-1a

AML-1aAML-1a AML1-aAML-1aAML-1a

AML-1aAML-1a

Oct-1 Sp-1

SRY SRY SRY SRY AP-1 AP-1 AP-1AP-1 AP-1 AP-1

GATA-2

p53

Figure 7.3. Illustration of putative transcription factors within the upstream BARD1 5’UTR region.

Page 142: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

133

After analysis of its sequence for possible transcription factors, we found several

sites of interest inferring that the –2402 region could be the regulatory region. This

stretch of DNA has to be evaluated in future studies. Unique NF-kB, p53 and p300

binding sites were found as well as a repetitive sites for C/EBP. The promoter is a

TATAless promoter region, rich in CpG and Sp-1 sites, which is characteristic of

constitutive promoters. However, the NF-κB site could be important in the

inflammation process and the regulation of BARD1 transcription. Further experiments

should include foot-printing, band shift as well as super shift assays in order to

determine the existence and the physical binding of the aforementioned transcription

factors on this stretch of DNA. Furthermore, since it is known that BRCA1 over-

expression can up-regulate NF-κB in vitro, it would be interesting to see if this region

can be activated by cytokines such as TNF-α or by BRCA1 transfection. This would

provide new insight on BARD1 regulation and the identification of its feedback loop.

BARD1 promoter response

100

1000

10000

100000

1000000

10000000

mock basic control (-) L5 L5 CE1transfected plasmids

luci

fera

se a

ctiv

ity TAC-2293 T3T3PC-3HeLa

Figure 7.4. Luciferase activity in 5 different cell lines. This histogram shows directional expression of –865 bp BARD1 minimal promoter region. Biological variation between different cell lines is probably due to transfection efficiency. Control is pGL3 control vector containing a firefly luciferase under a CMV promoter and an enhancer for optimal activity. (-)L5 is the 856 sequence from L5 and inserted negatively and serves as a mock. It demonstrates that orientation is primordial in this context despite several putative transcription factor sites placed in the opposite direction. Non specific DNA in pGL3 basic vector is immune to aberrant cryptic responsiveness. TAC-2 and NIH 3T3 are mouse cells the others are human. The human promoter region appears to show some conservation in mice.

Page 143: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

134

Furthermore, we know that BARD1 may bind to the p50 sub-unit of NF-κB, through

Bcl-3, and this could be the link in a possible regulated feedback loop for BARD1

expression.

293 T cells BARD1 promoter assay in duplicate

0

500000

1000000

1500000

2000000

2500000

mock basic control (-) L5 L5 CE1

Transfected samples

Luci

fera

se A

ctiv

ity

non-treatedDoxorubicin

3T3 BARD1 promoter assay duplicate analysis

0

100000

200000

300000

400000

500000

mock basic control neg L5 CE1

transfected plasmids

luci

fera

se a

ctiv

ity

non-treatedDoxorubicin

TAC-2 BARD1 promoter assay duplicate

0

10000002000000

30000004000000

5000000

mock basic control (-) L5 L5 CE1

transfected plasmids

luci

fera

se a

ctiv

ity

non-treatedDoxorubicin

PC-3 BARD1 promoter duplicate analysis

0

10000

20000

30000

40000

50000

60000

Non Trans empty vectorpGL3 basic

positive pGL3control

(-) L5 BARD1-pro

L5 BARD1-pro

CE1 BARD1-pro

transfected plasmids

luci

fera

se a

ctiv

ity

non-treatedDoxorubicin

HeLA BARD1 promoter assay (overall A+B)

0

1000

2000

3000

4000

5000

mock basic control (-) L5 L5 CE1

transfected plasmids

luci

fera

se a

ctiv

ity

Série1Doxorubicin

Figure 7.5. Dual luciferase assay after treatment with doxorubicin. No significant change in luciferase expression after treatment with 5 µg/ml for 6 hrs in any cell line used.

Page 144: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

135

Relative luciferase expression normalised to Renilla luciferase (TK-promoter)

050

100

150200250

pGL3control

pGL3basic

pGL3 L5(-) neg

sense

pGL3 L5

pGL3 CE1

plasmids transfected

Rel

ativ

e lu

cife

rase

uni

ts

in %

non-treatedHCGLHestrogendoxorubicin

Figure 7.6. Response of luciferase reporter gene after hormonal treatment. Dual luciferase assay.

Page 145: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

136

Conclusion This thesis is dispersed over several points but concentrate on the same theme: the

role of BARD1 in tumour suppression and maintenance of genomic stability through

cell cycle mechanisms. We wanted to investigate the role of BARD1 in tumour

suppression and how it controls genetic integrity. We have shown that BARD1 often

reflects the actions normally performed by BRCA1. Indeed BARD1 repression

creates genomic instability, loss of regular or normal cell cycle progression,

proliferation defects, and a tumour phenotype. Moreover, knockouts of BARD1,

BRCA1 or BRCA2 are not viable, since embryos die early in gestation. In this thesis

we have shown that BARD1 may have BRCA1-independent functions in addition of a

BRCA1-dependent role, such as pro-apoptotic response to genotoxic stress which is

dependent on p53. Furthermore, BARD1 is up-regulated upon genotoxic insult,

hypoxia and cell death in vitro and in vivo. BARD1 pro-apoptotic function has been

mapped to a region spanning amino acids 510 to 604. Moreover, this is confirmed by

the fact that the mutant Q564H is inefficient in promoting cell death. We have shown

that BARD1 is a dynamic protein with a function in the cytoplasm, expressed in G2/M

and which is associated with apoptosis. Furthermore, we showed that BARD1 binds

to p53 and Ku-70 in co-immunoprecipitation assays, independently from BRCA1.

BARD1 has many roles and could be mediated through BRCA1/BARD1 E3 ubiquitin

ligase activity. Lastly, it appears that BARD1 has several important roles in

spermatogenesis, embryogenesis, cell cycle progression, mitosis, centrosome

duplication and cytokinesis. It interacts with BRCA2 during G2/M and it also interacts

with TACC1 and Aurora B but not with Aurora A. In the future, it would be necessary

to dissect the precise role for BARD1 in this part of cell physiology. BARD1

commands vital cellular functions and is definitely a multifunctional protein that still

holds many secrets.

Page 146: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

137

References

1. Casey, G. The BRCA1 and BRCA2 breast cancer genes. Curr Opin Oncol 9, 88-93 (1997).

2. Wooster, R. & Weber, B. L. Breast and ovarian cancer. N Engl J Med 348, 2339-47 (2003).

3. Hall, J. M. et al. Linkage of early-onset familial breast cancer to chromosome 17q21. Science 250, 1684-9 (1990).

4. Futreal, P. A. et al. BRCA1 mutations in primary breast and ovarian carcinomas. Science 266, 120-2 (1994).

5. Miki, Y. et al. A strong candidate for the breast and ovarian cancer susceptibility gene BRCA1. Science 266, 66-71 (1994).

6. Wooster, R. et al. Localization of a breast cancer susceptibility gene, BRCA2, to chromosome 13q12-13. Science 265, 2088-90 (1994).

7. Wooster, R. et al. Identification of the breast cancer susceptibility gene BRCA2. Nature 378, 789-92 (1995).

8. Shugart, Y. Y. et al. Linkage analysis of 56 multiplex families excludes the Cowden disease gene PTEN as a major contributor to familial breast cancer. J Med Genet 36, 720-1 (1999).

9. Meijers-Heijboer, H. et al. Low-penetrance susceptibility to breast cancer due to CHEK2(*)1100delC in noncarriers of BRCA1 or BRCA2 mutations. Nat Genet 31, 55-9 (2002).

10. Schutte, M. et al. Variants in CHEK2 other than 1100delC do not make a major contribution to breast cancer susceptibility. Am J Hum Genet 72, 1023-8 (2003).

11. Caligo, M. A. et al. The CHEK2 c.1100delC mutation plays an irrelevant role in breast cancer predisposition in Italy. Hum Mutat 24, 100-1 (2004).

12. Smith, S. A. & Ponder, B. A. Predisposing genes in breast and ovarian cancer: an overview. Tumori 79, 291-6 (1993).

13. Ford, D. & Easton, D. F. The genetics of breast and ovarian cancer. Br J Cancer 72, 805-12 (1995).

14. Chen, J. & Lindblom, A. Germline mutation screening of the STK11/LKB1 gene in familial breast cancer with LOH on 19p. Clin Genet 57, 394-7 (2000).

15. Mincey, B. A. Genetics and the management of women at high risk for breast cancer. Oncologist 8, 466-73 (2003).

16. Easton, D., Ford, D. & Peto, J. Inherited susceptibility to breast cancer. Cancer Surv 18, 95-113 (1993).

17. Easton, D. F., Narod, S. A., Ford, D. & Steel, M. The genetic epidemiology of BRCA1. Breast Cancer Linkage Consortium. Lancet 344, 761 (1994).

18. Ford, D., Easton, D. F., Bishop, D. T., Narod, S. A. & Goldgar, D. E. Risks of cancer in BRCA1-mutation carriers. Breast Cancer Linkage Consortium. Lancet 343, 692-5 (1994).

19. Rahman, N. & Stratton, M. R. The genetics of breast cancer susceptibility. Annu Rev Genet 32, 95-121 (1998).

20. Ford, D. et al. Genetic heterogeneity and penetrance analysis of the BRCA1 and BRCA2 genes in breast cancer families. The Breast Cancer Linkage Consortium. Am J Hum Genet 62, 676-89 (1998).

Page 147: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

138

21. Holt, J. T. et al. Growth retardation and tumour inhibition by BRCA1. Nat Genet 12, 298-302 (1996).

22. Hakem, R. et al. The tumor suppressor gene Brca1 is required for embryonic cellular proliferation in the mouse. Cell 85, 1009-23 (1996).

23. Gowen, L. C., Johnson, B. L., Latour, A. M., Sulik, K. K. & Koller, B. H. Brca1 deficiency results in early embryonic lethality characterized by neuroepithelial abnormalities. Nat Genet 12, 191-4 (1996).

24. Monteiro, A. N. BRCA1: the enigma of tissue-specific tumor development. Trends Genet 19, 312-5 (2003).

25. Starita, L. M. & Parvin, J. D. The multiple nuclear functions of BRCA1: transcription, ubiquitination and DNA repair. Curr Opin Cell Biol 15, 345-50 (2003).

26. Venkitaraman, A. R. Cancer susceptibility and the functions of BRCA1 and BRCA2. Cell 108, 171-82 (2002).

27. Venkitaraman, A. R. Tracing the network connecting BRCA and Fanconi anaemia proteins. Nat Rev Cancer 4, 266-76 (2004).

28. Hakem, R., de la Pompa, J. L. & Mak, T. W. Developmental studies of Brca1 and Brca2 knock-out mice. J Mammary Gland Biol Neoplasia 3, 431-45 (1998).

29. Xu, X. et al. Conditional mutation of Brca1 in mammary epithelial cells results in blunted ductal morphogenesis and tumour formation. Nat Genet 22, 37-43 (1999).

30. Zheng, L., Li, S., Boyer, T. G. & Lee, W. H. Lessons learned from BRCA1 and BRCA2. Oncogene 19, 6159-75 (2000).

31. Scully, R. & Livingston, D. M. In search of the tumour-suppressor functions of BRCA1 and BRCA2. Nature 408, 429-32 (2000).

32. Wang, Q., Zhang, H., Fishel, R. & Greene, M. I. BRCA1 and cell signaling. Oncogene 19, 6152-8 (2000).

33. Kerr, P. & Ashworth, A. New complexities for BRCA1 and BRCA2. Curr Biol 11, R668-76 (2001).

34. Mueller, C. R. & Roskelley, C. D. Regulation of BRCA1 expression and its relationship to sporadic breast cancer. Breast Cancer Res 5, 45-52 (2003).

35. Moynahan, M. E. The cancer connection: BRCA1 and BRCA2 tumor suppression in mice and humans. Oncogene 21, 8994-9007 (2002).

36. Narod, S. A. & Foulkes, W. D. BRCA1 and BRCA2: 1994 and beyond. Nat Rev Cancer 4, 665-76 (2004).

37. Deng, C. X. & Brodie, S. G. Roles of BRCA1 and its interacting proteins. Bioessays 22, 728-37 (2000).

38. Khanna, K. K. & Jackson, S. P. DNA double-strand breaks: signaling, repair and the cancer connection. Nat Genet 27, 247-54 (2001).

39. Hsu, L. C. & White, R. L. BRCA1 is associated with the centrosome during mitosis. Proc Natl Acad Sci U S A 95, 12983-8 (1998).

40. Hsu, L. C., Doan, T. P. & White, R. L. Identification of a gamma-tubulin-binding domain in BRCA1. Cancer Res 61, 7713-8 (2001).

41. Schlegel, B. P., Starita, L. M. & Parvin, J. D. Overexpression of a protein fragment of RNA helicase A causes inhibition of endogenous BRCA1 function and defects in ploidy and cytokinesis in mammary epithelial cells. Oncogene 22, 983-91 (2003).

42. Rajan, J. V., Marquis, S. T., Gardner, H. P. & Chodosh, L. A. Developmental expression of Brca2 colocalizes with Brca1 and is associated with proliferation and differentiation in multiple tissues. Dev Biol 184, 385-401 (1997).

43. Jasin, M. Homologous repair of DNA damage and tumorigenesis: the BRCA connection. Oncogene 21, 8981-93 (2002).

Page 148: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

139

44. Wu, L. C. et al. Identification of a RING protein that can interact in vivo with the BRCA1 gene product. Nat Genet 14, 430-40 (1996).

45. Thai, T. H. et al. Mutations in the BRCA1-associated RING domain (BARD1) gene in primary breast, ovarian and uterine cancers. Hum Mol Genet 7, 195-202 (1998).

46. Connor, F. et al. Cloning, chromosomal mapping and expression pattern of the mouse Brca2 gene. Hum Mol Genet 6, 291-300 (1997).

47. Irminger-Finger, I., Soriano, J. V., Vaudan, G., Montesano, R. & Sappino, A. P. In vitro repression of Brca1-associated RING domain gene, Bard1, induces phenotypic changes in mammary epithelial cells. J Cell Biol 143, 1329-39 (1998).

48. Ayi, T. C., Tsan, J. T., Hwang, L. Y., Bowcock, A. M. & Baer, R. Conservation of function and primary structure in the BRCA1-associated RING domain (BARD1) protein. Oncogene 17, 2143-8 (1998).

49. Joukov, V., Chen, J., Fox, E. A., Green, J. B. & Livingston, D. M. Functional communication between endogenous BRCA1 and its partner, BARD1, during Xenopus laevis development. Proc Natl Acad Sci U S A 98, 12078-83 (2001).

50. Boulton, S. J. et al. BRCA1/BARD1 orthologs required for DNA repair in Caenorhabditis elegans. Curr Biol 14, 33-9 (2004).

51. Jefford, C. E., Feki, A., Harb, J., Krause, K. H. & Irminger-Finger, I. Nuclear-cytoplasmic translocation of BARD1 is linked to its apoptotic activity. Oncogene 23, 3509-20 (2004).

52. Borden, K. L. RING fingers and B-boxes: zinc-binding protein-protein interaction domains. Biochem Cell Biol 76, 351-8 (1998).

53. Freemont, P. S. RING for destruction? Curr Biol 10, R84-7 (2000). 54. Kentsis, A. & Borden, K. L. Physical mechanisms and biological significance of

supramolecular protein self-assembly. Curr Protein Pept Sci 5, 125-34 (2004). 55. Meza, J. E., Brzovic, P. S., King, M. C. & Klevit, R. E. Mapping the functional

domains of BRCA1. Interaction of the ring finger domains of BRCA1 and BARD1. J Biol Chem 274, 5659-65 (1999).

56. Brzovic, P. S., Rajagopal, P., Hoyt, D. W., King, M. C. & Klevit, R. E. Structure of a BRCA1-BARD1 heterodimeric RING-RING complex. Nat Struct Biol 8, 833-7 (2001).

57. Morris, J. R., Keep, N. H. & Solomon, E. Identification of residues required for the interaction of BARD1 with BRCA1. J Biol Chem 277, 9382-6 (2002).

58. Brzovic, P. S., Meza, J. E., King, M. C. & Klevit, R. E. BRCA1 RING domain cancer-predisposing mutations. Structural consequences and effects on protein-protein interactions. J Biol Chem 276, 41399-406 (2001).

59. Brzovic, P. S. et al. Binding and recognition in the assembly of an active BRCA1/BARD1 ubiquitin-ligase complex. Proc Natl Acad Sci U S A 100, 5646-51 (2003).

60. Mosavi, L. K., Cammett, T. J., Desrosiers, D. C. & Peng, Z. Y. The ankyrin repeat as molecular architecture for protein recognition. Protein Sci 13, 1435-48 (2004).

61. Zhang, X. et al. Structure of an XRCC1 BRCT domain: a new protein-protein interaction module. Embo J 17, 6404-11 (1998).

62. Huyton, T., Bates, P. A., Zhang, X., Sternberg, M. J. & Freemont, P. S. The BRCA1 C-terminal domain: structure and function. Mutat Res 460, 319-32 (2000).

63. Williams, R. S., Green, R. & Glover, J. N. Crystal structure of the BRCT repeat region from the breast cancer-associated protein BRCA1. Nat Struct Biol 8, 838-42 (2001).

64. Ljungquist, S., Kenne, K., Olsson, L. & Sandstrom, M. Altered DNA ligase III activity in the CHO EM9 mutant. Mutat Res 314, 177-86 (1994).

Page 149: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

140

65. Caldecott, K. W., Tucker, J. D., Stanker, L. H. & Thompson, L. H. Characterization of the XRCC1-DNA ligase III complex in vitro and its absence from mutant hamster cells. Nucleic Acids Res 23, 4836-43 (1995).

66. Friedman, L. S. et al. Confirmation of BRCA1 by analysis of germline mutations linked to breast and ovarian cancer in ten families. Nat Genet 8, 399-404 (1994).

67. Cantor, S. B. et al. BACH1, a novel helicase-like protein, interacts directly with BRCA1 and contributes to its DNA repair function. Cell 105, 149-60 (2001).

68. Ohta, T. & Fukuda, M. Ubiquitin and breast cancer. Oncogene 23, 2079-88 (2004). 69. Weissman, A. M. Themes and variations on ubiquitylation. Nat Rev Mol Cell Biol 2,

169-78 (2001). 70. Pickart, C. M. Ubiquitin in chains. Trends Biochem Sci 25, 544-8 (2000). 71. Hicke, L. Protein regulation by monoubiquitin. Nat Rev Mol Cell Biol 2, 195-201

(2001). 72. Hershko, A. & Ciechanover, A. The ubiquitin system. Annu Rev Biochem 67, 425-79

(1998). 73. Wilkinson, K. D. Ubiquitination and deubiquitination: targeting of proteins for

degradation by the proteasome. Semin Cell Dev Biol 11, 141-8 (2000). 74. Huibregtse, J. M., Scheffner, M., Beaudenon, S. & Howley, P. M. A family of proteins

structurally and functionally related to the E6-AP ubiquitin-protein ligase. Proc Natl Acad Sci U S A 92, 2563-7 (1995).

75. Joazeiro, C. A. & Weissman, A. M. RING finger proteins: mediators of ubiquitin ligase activity. Cell 102, 549-52 (2000).

76. Hashizume, R. et al. The RING heterodimer BRCA1-BARD1 is a ubiquitin ligase inactivated by a breast cancer-derived mutation. J Biol Chem 276, 14537-40 (2001).

77. Ruffner, H., Joazeiro, C. A., Hemmati, D., Hunter, T. & Verma, I. M. Cancer-predisposing mutations within the RING domain of BRCA1: loss of ubiquitin protein ligase activity and protection from radiation hypersensitivity. Proc Natl Acad Sci U S A 98, 5134-9 (2001).

78. Xia, Y., Pao, G. M., Chen, H. W., Verma, I. M. & Hunter, T. Enhancement of BRCA1 E3 ubiquitin ligase activity through direct interaction with the BARD1 protein. J Biol Chem 278, 5255-63 (2003).

79. Kentsis, A., Gordon, R. E. & Borden, K. L. Control of biochemical reactions through supramolecular RING domain self-assembly. Proc Natl Acad Sci U S A 99, 15404-9 (2002).

80. Chen, A., Kleiman, F. E., Manley, J. L., Ouchi, T. & Pan, Z. Q. Autoubiquitination of the BRCA1*BARD1 RING ubiquitin ligase. J Biol Chem 277, 22085-92 (2002).

81. Wu-Baer, F., Lagrazon, K., Yuan, W. & Baer, R. The BRCA1/BARD1 heterodimer assembles polyubiquitin chains through an unconventional linkage involving lysine residue K6 of ubiquitin. J Biol Chem 278, 34743-6 (2003).

82. Nishikawa, H. et al. Mass spectrometric and mutational analyses reveal Lys-6-linked polyubiquitin chains catalyzed by BRCA1-BARD1 ubiquitin ligase. J Biol Chem 279, 3916-24 (2004).

83. Morris, J. R. & Solomon, E. BRCA1 : BARD1 induces the formation of conjugated ubiquitin structures, dependent on K6 of ubiquitin, in cells during DNA replication and repair. Hum Mol Genet 13, 807-17 (2004).

84. Mallery, D. L., Vandenberg, C. J. & Hiom, K. Activation of the E3 ligase function of the BRCA1/BARD1 complex by polyubiquitin chains. Embo J 21, 6755-62 (2002).

85. Choudhury, A. D., Xu, H. & Baer, R. Ubiquitination and proteasomal degradation of the BRCA1 tumor suppressor is regulated during cell cycle progression. J Biol Chem 279, 33909-18 (2004).

Page 150: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

141

86. Levinger, L. & Varshavsky, A. Selective arrangement of ubiquitinated and D1 protein-containing nucleosomes within the Drosophila genome. Cell 28, 375-85 (1982).

87. Davie, J. R. & Murphy, L. C. Level of ubiquitinated histone H2B in chromatin is coupled to ongoing transcription. Biochemistry 29, 4752-7 (1990).

88. Davie, J. R., Lin, R. & Allis, C. D. Timing of the appearance of ubiquitinated histones in developing new macronuclei of Tetrahymena thermophila. Biochem Cell Biol 69, 66-71 (1991).

89. Garcia-Higuera, I. et al. Interaction of the Fanconi anemia proteins and BRCA1 in a common pathway. Mol Cell 7, 249-62 (2001).

90. Vandenberg, C. J. et al. BRCA1-independent ubiquitination of FANCD2. Mol Cell 12, 247-54 (2003).

91. Starita, L. M. et al. BRCA1-dependent ubiquitination of gamma-tubulin regulates centrosome number. Mol Cell Biol 24, 8457-66 (2004).

92. Sato, K. et al. Nucleophosmin/B23 is a candidate substrate for the BRCA1-BARD1 ubiquitin ligase. J Biol Chem 279, 30919-22 (2004).

93. Hayami, R. et al. Down-regulation of BRCA1-BARD1 ubiquitin ligase by CDK2. Cancer Res 65, 6-10 (2005).

94. Dong, Y. et al. Regulation of BRCC, a holoenzyme complex containing BRCA1 and BRCA2, by a signalosome-like subunit and its role in DNA repair. Mol Cell 12, 1087-99 (2003).

95. Jensen, D. E. et al. BAP1: a novel ubiquitin hydrolase which binds to the BRCA1 RING finger and enhances BRCA1-mediated cell growth suppression. Oncogene 16, 1097-112 (1998).

96. Liu, C. Y., Flesken-Nikitin, A., Li, S., Zeng, Y. & Lee, W. H. Inactivation of the mouse Brca1 gene leads to failure in the morphogenesis of the egg cylinder in early postimplantation development. Genes Dev 10, 1835-43 (1996).

97. Ludwig, T., Chapman, D. L., Papaioannou, V. E. & Efstratiadis, A. Targeted mutations of breast cancer susceptibility gene homologs in mice: lethal phenotypes of Brca1, Brca2, Brca1/Brca2, Brca1/p53, and Brca2/p53 nullizygous embryos. Genes Dev 11, 1226-41 (1997).

98. McCarthy, E. E., Celebi, J. T., Baer, R. & Ludwig, T. Loss of Bard1, the heterodimeric partner of the Brca1 tumor suppressor, results in early embryonic lethality and chromosomal instability. Mol Cell Biol 23, 5056-63 (2003).

99. Gowen, L. C., Avrutskaya, A. V., Latour, A. M., Koller, B. H. & Leadon, S. A. BRCA1 required for transcription-coupled repair of oxidative DNA damage. Science 281, 1009-12 (1998).

100. Moynahan, M. E., Chiu, J. W., Koller, B. H. & Jasin, M. Brca1 controls homology-directed DNA repair. Mol Cell 4, 511-8 (1999).

101. Scully, R. et al. Genetic analysis of BRCA1 function in a defined tumor cell line. Mol Cell 4, 1093-9 (1999).

102. Abbott, D. W. et al. BRCA1 expression restores radiation resistance in BRCA1-defective cancer cells through enhancement of transcription-coupled DNA repair. J Biol Chem 274, 18808-12 (1999).

103. Snouwaert, J. N. et al. BRCA1 deficient embryonic stem cells display a decreased homologous recombination frequency and an increased frequency of non-homologous recombination that is corrected by expression of a brca1 transgene. Oncogene 18, 7900-7 (1999).

104. Zhong, Q., Boyer, T. G., Chen, P. L. & Lee, W. H. Deficient nonhomologous end-joining activity in cell-free extracts from Brca1-null fibroblasts. Cancer Res 62, 3966-70 (2002).

Page 151: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

142

105. Zhong, Q., Chen, C. F., Chen, P. L. & Lee, W. H. BRCA1 facilitates microhomology-mediated end joining of DNA double strand breaks. J Biol Chem 277, 28641-7 (2002).

106. Westermark, U. K. et al. BARD1 participates with BRCA1 in homology-directed repair of chromosome breaks. Mol Cell Biol 23, 7926-36 (2003).

107. Wang, Q. et al. Adenosine nucleotide modulates the physical interaction between hMSH2 and BRCA1. Oncogene 20, 4640-9 (2001).

108. Jin, Y. et al. Cell cycle-dependent colocalization of BARD1 and BRCA1 proteins in discrete nuclear domains. Proc Natl Acad Sci U S A 94, 12075-80 (1997).

109. Chen, Y. et al. BRCA1 is a 220-kDa nuclear phosphoprotein that is expressed and phosphorylated in a cell cycle-dependent manner. Cancer Res 56, 3168-72 (1996).

110. Vaughn, J. P. et al. BRCA1 expression is induced before DNA synthesis in both normal and tumor-derived breast cells. Cell Growth Differ 7, 711-5 (1996).

111. Gudas, J. M. et al. Cell cycle regulation of BRCA1 messenger RNA in human breast epithelial cells. Cell Growth Differ 7, 717-23 (1996).

112. Rajan, J. V., Wang, M., Marquis, S. T. & Chodosh, L. A. Brca2 is coordinately regulated with Brca1 during proliferation and differentiation in mammary epithelial cells. Proc Natl Acad Sci U S A 93, 13078-83 (1996).

113. Scully, R. et al. Dynamic changes of BRCA1 subnuclear location and phosphorylation state are initiated by DNA damage. Cell 90, 425-35 (1997).

114. Paull, T. T., Cortez, D., Bowers, B., Elledge, S. J. & Gellert, M. Direct DNA binding by Brca1. Proc Natl Acad Sci U S A 98, 6086-91 (2001).

115. Chiba, N. & Parvin, J. D. Redistribution of BRCA1 among four different protein complexes following replication blockage. J Biol Chem 276, 38549-54 (2001).

116. Zhong, Q. et al. Association of BRCA1 with the hRad50-hMre11-p95 complex and the DNA damage response. Science 285, 747-50 (1999).

117. Chapman, M. S. & Verma, I. M. Transcriptional activation by BRCA1. Nature 382, 678-9 (1996).

118. Monteiro, A. N., August, A. & Hanafusa, H. Evidence for a transcriptional activation function of BRCA1 C-terminal region. Proc Natl Acad Sci U S A 93, 13595-9 (1996).

119. Monteiro, A. N., August, A. & Hanafusa, H. Common BRCA1 variants and transcriptional activation. Am J Hum Genet 61, 761-2 (1997).

120. Haile, D. T. & Parvin, J. D. Activation of transcription in vitro by the BRCA1 carboxyl-terminal domain. J Biol Chem 274, 2113-7 (1999).

121. Anderson, S. F., Schlegel, B. P., Nakajima, T., Wolpin, E. S. & Parvin, J. D. BRCA1 protein is linked to the RNA polymerase II holoenzyme complex via RNA helicase A. Nat Genet 19, 254-6 (1998).

122. Scully, R. et al. BRCA1 is a component of the RNA polymerase II holoenzyme. Proc Natl Acad Sci U S A 94, 5605-10 (1997).

123. Neish, A. S., Anderson, S. F., Schlegel, B. P., Wei, W. & Parvin, J. D. Factors associated with the mammalian RNA polymerase II holoenzyme. Nucleic Acids Res 26, 847-53 (1998).

124. Chiba, N. & Parvin, J. D. The BRCA1 and BARD1 association with the RNA polymerase II holoenzyme. Cancer Res 62, 4222-8 (2002).

125. Baldwin, A. S., Jr. The NF-kappa B and I kappa B proteins: new discoveries and insights. Annu Rev Immunol 14, 649-83 (1996).

126. Baldwin, A. S. Control of oncogenesis and cancer therapy resistance by the transcription factor NF-kappaB. J Clin Invest 107, 241-6 (2001).

127. Dechend, R. et al. The Bcl-3 oncoprotein acts as a bridging factor between NF-kappaB/Rel and nuclear co-regulators. Oncogene 18, 3316-23 (1999).

Page 152: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

143

128. Benezra, M. et al. BRCA1 augments transcription by the NF-kappaB transcription factor by binding to the Rel domain of the p65/RelA subunit. J Biol Chem 278, 26333-41 (2003).

129. Chen, L. F. & Greene, W. C. Regulation of distinct biological activities of the NF-kappaB transcription factor complex by acetylation. J Mol Med 81, 549-57 (2003).

130. Cui, J. Q. et al. BRCA1 splice variants BRCA1a and BRCA1b associate with CBP co-activator. Oncol Rep 5, 591-5 (1998).

131. Pao, G. M., Janknecht, R., Ruffner, H., Hunter, T. & Verma, I. M. CBP/p300 interact with and function as transcriptional coactivators of BRCA1. Proc Natl Acad Sci U S A 97, 1020-5 (2000).

132. von Mikecz, A., Zhang, S., Montminy, M., Tan, E. M. & Hemmerich, P. CREB-binding protein (CBP)/p300 and RNA polymerase II colocalize in transcriptionally active domains in the nucleus. J Cell Biol 150, 265-73 (2000).

133. Ogryzko, V. V., Schiltz, R. L., Russanova, V., Howard, B. H. & Nakatani, Y. The transcriptional coactivators p300 and CBP are histone acetyltransferases. Cell 87, 953-9 (1996).

134. Bannister, A. J. & Kouzarides, T. The CBP co-activator is a histone acetyltransferase. Nature 384, 641-3 (1996).

135. Bochar, D. A. et al. BRCA1 is associated with a human SWI/SNF-related complex: linking chromatin remodeling to breast cancer. Cell 102, 257-65 (2000).

136. Ganesan, S. et al. BRCA1 supports XIST RNA concentration on the inactive X chromosome. Cell 111, 393-405 (2002).

137. Scully, R. et al. Association of BRCA1 with Rad51 in mitotic and meiotic cells. Cell 88, 265-75 (1997).

138. Kleiman, F. E. & Manley, J. L. Functional interaction of BRCA1-associated BARD1 with polyadenylation factor CstF-50. Science 285, 1576-9 (1999).

139. Takagaki, Y., Manley, J. L., MacDonald, C. C., Wilusz, J. & Shenk, T. A multisubunit factor, CstF, is required for polyadenylation of mammalian pre-mRNAs. Genes Dev 4, 2112-20 (1990).

140. Takagaki, Y. & Manley, J. L. RNA recognition by the human polyadenylation factor CstF. Mol Cell Biol 17, 3907-14 (1997).

141. Kleiman, F. E. & Manley, J. L. The BARD1-CstF-50 interaction links mRNA 3' end formation to DNA damage and tumor suppression. Cell 104, 743-53 (2001).

142. Colgan, D. F. & Manley, J. L. Mechanism and regulation of mRNA polyadenylation. Genes Dev 11, 2755-66 (1997).

143. Zhao, W. & Manley, J. L. Deregulation of poly(A) polymerase interferes with cell growth. Mol Cell Biol 18, 5010-20 (1998).

144. Spahn, L. et al. Interaction of the EWS NH2 terminus with BARD1 links the Ewing's sarcoma gene to a common tumor suppressor pathway. Cancer Res 62, 4583-7 (2002).

145. Bruun, D. et al. siRNA depletion of BRCA1, but not BRCA2, causes increased genome instability in Fanconi anemia cells. DNA Repair (Amst) 2, 1007-13 (2003).

146. Irminger-Finger, I. et al. Identification of BARD1 as mediator between proapoptotic stress and p53-dependent apoptosis. Mol Cell 8, 1255-66 (2001).

147. Meek, D. W. Post-translational modification of p53. Semin Cancer Biol 5, 203-10 (1994).

148. Milczarek, G. J., Martinez, J. & Bowden, G. T. p53 Phosphorylation: biochemical and functional consequences. Life Sci 60, 1-11 (1997).

149. Fabbro, M. et al. BRCA1-BARD1 complexes are required for p53Ser-15 phosphorylation and a G1/S arrest following ionizing radiation-induced DNA damage. J Biol Chem 279, 31251-8 (2004).

Page 153: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

144

150. Feki, A. et al. BARD1 induces apoptosis by catalysing phosphorylation of p53 by DNA-damage response kinase. Oncogene (2005).

151. Chai, Y. L. et al. The second BRCT domain of BRCA1 proteins interacts with p53 and stimulates transcription from the p21WAF1/CIP1 promoter. Oncogene 18, 263-8 (1999).

152. Ouchi, T., Monteiro, A. N., August, A., Aaronson, S. A. & Hanafusa, H. BRCA1 regulates p53-dependent gene expression. Proc Natl Acad Sci U S A 95, 2302-6 (1998).

153. Arizti, P. et al. Tumor suppressor p53 is required to modulate BRCA1 expression. Mol Cell Biol 20, 7450-9 (2000).

154. Harkin, D. P. et al. Induction of GADD45 and JNK/SAPK-dependent apoptosis following inducible expression of BRCA1. Cell 97, 575-86 (1999).

155. Jensen, R. A. et al. BRCA1 is secreted and exhibits properties of a granin. Nat Genet 12, 303-8 (1996).

156. Rodriguez, J. A. & Henderson, B. R. Identification of a functional nuclear export sequence in BRCA1. J Biol Chem 275, 38589-96 (2000).

157. Fabbro, M., Rodriguez, J. A., Baer, R. & Henderson, B. R. BARD1 induces BRCA1 intranuclear foci formation by increasing RING-dependent BRCA1 nuclear import and inhibiting BRCA1 nuclear export. J Biol Chem 277, 21315-24 (2002).

158. Schuechner, S., Tembe, V., Rodriguez, J. A. & Henderson, B. R. Nuclear targeting and cell cycle regulatory function of human BARD1. J Biol Chem (2005).

159. Rodriguez, J. A., Schuchner, S., Au, W. W., Fabbro, M. & Henderson, B. R. Nuclear-cytoplasmic shuttling of BARD1 contributes to its proapoptotic activity and is regulated by dimerization with BRCA1. Oncogene 23, 1809-20 (2004).

160. Ghimenti, C. et al. Germline mutations of the BRCA1-associated ring domain (BARD1) gene in breast and breast/ovarian families negative for BRCA1 and BRCA2 alterations. Genes Chromosomes Cancer 33, 235-42 (2002).

161. Karppinen, S. M., Heikkinen, K., Rapakko, K. & Winqvist, R. Mutation screening of the BARD1 gene: evidence for involvement of the Cys557Ser allele in hereditary susceptibility to breast cancer. J Med Genet 41, e114 (2004).

162. Gautier, F., Irminger-Finger, I., Gregoire, M., Meflah, K. & Harb, J. Identification of an apoptotic cleavage product of BARD1 as an autoantigen: a potential factor in the antitumoral response mediated by apoptotic bodies. Cancer Res 60, 6895-900 (2000).

163. Daniels, M. J., Wang, Y., Lee, M. & Venkitaraman, A. R. Abnormal cytokinesis in cells deficient in the breast cancer susceptibility protein BRCA2. Science 306, 876-9 (2004).

164. Brzovic, P. S., Meza, J., King, M. C. & Klevit, R. E. The cancer-predisposing mutation C61G disrupts homodimer formation in the NH2-terminal BRCA1 RING finger domain. J Biol Chem 273, 7795-9 (1998).

165. Ishitobi, M. et al. Mutational analysis of BARD1 in familial breast cancer patients in Japan. Cancer Lett 200, 1-7 (2003).

166. Orban, T. I. & Olah, E. Expression profiles of BRCA1 splice variants in asynchronous and in G1/S synchronized tumor cell lines. Biochem Biophys Res Commun 280, 32-8 (2001).

167. McEachern, K. A., Archey, W. B., Douville, K. & Arrick, B. A. BRCA1 splice variants exhibit overlapping and distinct transcriptional transactivation activities. J Cell Biochem 89, 120-32 (2003).

168. Feki, A. & Irminger-Finger, I. Mutational spectrum of p53 mutations in primary breast and ovarian tumors. Crit Rev Oncol Hematol 52, 103-16 (2004).

Page 154: Mechanisms of tumour suppression and control › record › 5337 › files › 1_these-JeffordCE.pdf · sciences during a fulfilling internship, as well as Mr J. S. Johnson who provided

145

169. Feki, A. et al. BARD1 expression during spermatogenesis is associated with apoptosis and hormonally regulated. Biol Reprod 71, 1614-24 (2004).

170. Thompson, M. E., Jensen, R. A., Obermiller, P. S., Page, D. L. & Holt, J. T. Decreased expression of BRCA1 accelerates growth and is often present during sporadic breast cancer progression. Nat Genet 9, 444-50 (1995).

171. Yoshikawa, K. et al. Abnormal expression of BRCA1 and BRCA1-interactive DNA-repair proteins in breast carcinomas. Int J Cancer 88, 28-36 (2000).

172. Reinholz, M. M. et al. Differential gene expression of TGF beta inducible early gene (TIEG), Smad7, Smad2 and Bard1 in normal and malignant breast tissue. Breast Cancer Res Treat 86, 75-88 (2004).

173. Qiu, Y., Langman, M. J. & Eggo, M. C. Targets of 17beta-oestradiol-induced apoptosis in colon cancer cells: a mechanism for the protective effects of hormone replacement therapy? J Endocrinol 181, 327-37 (2004).