Download Hereditary predisposition to breast cancer – with a focus on AATF

Document related concepts
Transcript
D 1247
OULU 2014
UNIV ER S IT Y OF OULU P. O. B R[ 00 FI-90014 UNIVERSITY OF OULU FINLAND
U N I V E R S I TAT I S
S E R I E S
SCIENTIAE RERUM NATURALIUM
Professor Esa Hohtola
HUMANIORA
University Lecturer Santeri Palviainen
TECHNICA
Postdoctoral research fellow Sanna Taskila
MEDICA
Professor Olli Vuolteenaho
SCIENTIAE RERUM SOCIALIUM
ACTA
U N I V E R S I T AT I S O U L U E N S I S
Maria Haanpää
E D I T O R S
Maria Haanpää
A
B
C
D
E
F
G
O U L U E N S I S
ACTA
A C TA
D 1247
HEREDITARY
PREDISPOSITION TO BREAST
CANCER – WITH A FOCUS
ON AATF, MRG15, PALB2, AND
THREE FANCONI ANAEMIA
GENES
University Lecturer Veli-Matti Ulvinen
SCRIPTA ACADEMICA
Director Sinikka Eskelinen
OECONOMICA
Professor Jari Juga
EDITOR IN CHIEF
Professor Olli Vuolteenaho
PUBLICATIONS EDITOR
Publications Editor Kirsti Nurkkala
ISBN 978-952-62-0457-4 (Paperback)
ISBN 978-952-62-0458-1 (PDF)
ISSN 0355-3221 (Print)
ISSN 1796-2234 (Online)
UNIVERSITY OF OULU GRADUATE SCHOOL;
UNIVERSITY OF OULU, FACULTY OF MEDICINE,
INSTITUTE OF CLINICAL MEDICINE, DEPARTMENT OF CLINICAL GENETICS;
FACULTY OF MEDICINE, INSTITUTE OF DIAGNOSTICS,
DEPARTMENT OF CLINICAL CHEMISTRY;
BIOCENTER OULU;
OULU UNIVERSITY HOSPITAL;
NATIONAL GRADUATE SCHOOL OF CLINICAL INVESTIGATION
D
MEDICA
ACTA UNIVERSITATIS OULUENSIS
D Medica 1247
MARIA HAANPÄÄ
HEREDITARY PREDISPOSITION TO
BREAST CANCER – WITH A FOCUS
ON AATF, MRG15, PALB2, AND THREE
FANCONI ANAEMIA GENES
Academic dissertation to be presented with the assent of
the Doctoral Training Committee of Health and
Biosciences of the University of Oulu for public defence in
Auditorium 4 of Oulu University Hospital, on 6 June
2014, at 12 noon
U N I VE R S I T Y O F O U L U , O U L U 2 0 1 4
Copyright © 2014
Acta Univ. Oul. D 1247, 2014
Supervised by
Professor Robert Winqvist
Docent Katri Pylkäs
Reviewed by
Professor Helena Kääriäinen
Docent Pia Vahteristo
Opponent
Docent Annika Auranen
ISBN 978-952-62-0457-4 (Paperback)
ISBN 978-952-62-0458-1 (PDF)
ISSN 0355-3221 (Printed)
ISSN 1796-2234 (Online)
Cover Design
Raimo Ahonen
JUVENES PRINT
TAMPERE 2014
Haanpää, Maria, Hereditary predisposition to breast cancer – with a focus on
AATF, MRG15, PALB2, and three Fanconi anaemia genes.
University of Oulu Graduate School; University of Oulu, Faculty of Medicine, Institute of
Clinical Medicine, Department of Clinical Genetics; Faculty of Medicine, Institute of
Diagnostics, Department of Clinical Chemistry; Biocenter Oulu; Oulu University Hospital;
National Graduate School of Clinical Investigation
Acta Univ. Oul. D 1247, 2014
University of Oulu, P.O. Box 8000, FI-90014 University of Oulu, Finland
Abstract
Around 5−10% of all breast cancer cases are estimated to result from a strong hereditary
predisposition to the disease. However, mutations in the currently known breast cancer
susceptibility genes account for only 20−30% of all familial cases. Additional factors contributing
to the pathogenesis of breast cancer, therefore, await discovery. Aims of this study were to
evaluate variations of the AATF and MRG15 genes as novel potential candidates for breast cancer
susceptibility, to further examine the prevalence of the cancer-related PALB2 c.1592delT mutation
among BRCA-negative high-risk breast cancer families counselled at the Department of Clinical
Genetics, Oulu University Hospital, to identify Finnish Fanconi anaemia patients
complementation groups as well as causative mutations, and to evaluate the potential role of these
mutations in breast cancer susceptibility.
The analysis of 121 familial breast cancer cases revealed altogether seven different sequence
changes in the AATF gene. However, none of them were considered pathogenic, suggesting that
germline mutations in AATF are rare or absent in breast cancer patients.
Investigation of the MRG15 gene among familial breast cancer cases revealed seven previously
unreported variants, but in silico analyses revealed that none of these variants appeared to modify
the function of MRG15. The results suggest that MRG15 alterations are unlikely to be significant
breast cancer susceptibility alleles.
A previously identified pathogenic PALB2 mutation, c.1592delT, was identified in three
patients from a cohort of 62 high-risk BRCA1/2-negative breast cancer patients from the
Department of Clinical Genetics. PALB2 c.1592delT mutation testing should thus be a routine part
of the genetic counselling protocol, particularly for BRCA1/2-negative high-risk breast cancer
patients.
Investigation of the complementation groups of Finnish Fanconi anaemia patients revealed a
total of six different causative mutations. These mutations were examined further by analysing
their prevalence in large cohorts of breast (n=1840) and prostate (n=565) cancers. However, no
significant association emerged between cancer predisposition and these FA mutations.
Keywords: AATF, breast cancer, DNA damage response, DNA repair, Fanconi
anaemia, genetic predisposition to disease, germline mutation, MRG15, PALB2
Haanpää, Maria, Perinnöllinen rintasyöpäalttius – AATF, MRG15, PALB2 ja
kolme Fanconin anemian geeniä tutkimuksen kohteena.
Oulun yliopiston tutkijakoulu; Oulun yliopisto, Lääketieteellinen tiedekunta, Kliinisen
lääketieteen laitos, Perinnöllisyyslääketiede; Lääketieteellinen tiedekunta, Diagnostiikan laitos,
Kliininen kemia; Biocenter Oulu; Oulun yliopistollinen sairaala; Valtakunnallinen kliininen
tutkijakoulu
Acta Univ. Oul. D 1247, 2014
Oulun yliopisto, PL 8000, 90014 Oulun yliopisto
Tiivistelmä
Arviolta 5−10 prosenttia kaikista rintasyöpätapauksista aiheutuu merkittävästä perinnöllisestä
alttiudesta sairauteen. Tällä hetkellä tiedossa olevien rintasyövälle altistavien geenivirheiden ajatellaan kuitenkin selittävän vain noin 20−30 prosenttia kaikista perinnöllisistä tapauksista. On
todennäköistä, että uusia tekijöitä, jotka osallistuvat rintasyövän patomekanismiin, on vielä löytymättä. Tämän tutkimuksen tarkoituksena oli arvioida AATF- ja MRG15-geeneissä esiintyvien
muutosten mahdollista vaikutusta rintasyöpäalttiuteen, tutkia tarkemmin PALB2 c.1592delT mutaation esiintymistä BRCA-mutaationegatiivisten korkean rintasyöpäriskin potilaiden joukossa (perinnöllisyyspoliklinikka, Oulun yliopistollinen sairaala) ja määrittää suomalaisten Fanconianemiapotilaiden komplementaatioryhmät, sairauden taustalla olevat mutaatiot sekä tutkia näihin mutaatioihin mahdollisesti liittyvää rintasyöpäriskiä.
121 familiaalisen rintasyöpätapauksen analyysissä löytyi yhteensä seitsemän erilaista sekvenssimuutosta AATF-geenissä. Näistä yksikään ei kuitenkaan ollut selvästi patogeeninen.
Tuloksen perusteella perinnölliset rintasyövälle altistavat muutokset AATF-geenissä ovat joko
erittäin harvinaisia tai niitä ei esiinny lainkaan.
MRG15-geenin mutaatioanalyysissä havaittiin seitsemän aikaisemmin raportoimatonta muutosta, mutta in silico -analyysien perusteella millään muutoksista ei ole vaikutusta MRG15-proteiinin toimintaan. Tulosten perusteella on epätodennäköistä, että MRG15-geenin muutokset olisivat merkittäviä rintasyövälle altistavia muutoksia.
Jo aiemmin patogeeniseksi todettu PALB2 c.1592delT -mutaatio löydettiin kolmelta niistä
perinnöllisyyspoliklinikan korkean syöpäriskin 62 potilaasta, jotka olivat BRCA1/2-geenitestauksessa saaneet normaalin tuloksen. Tulostemme perusteella PALB2 c.1592delT -mutaatiotestaus tulisi Suomessa ottaa osaksi perinnöllisyyspoliklinikoiden tarjomaa tutkimusprotokollaa.
Suomalaisten Fanconi-anemiapotilaiden komplementaatioryhmiä selvittävässä tutkimuksessa identifioitiin yhteensä kuusi erilaista tautia aiheuttavaa mutaatiota. Näiden muutosten esiintymistä tutkittiin myös laajoissa rinta- (n=1840) ja eturauhassyöpäaineistoissa (n=565). Tilastollisesti merkittävää assosiaatiota ei kuitenkaan todettu suomalaisten FA-mutaatioiden ja syöpäalttiuden välillä.
Asiasanat: AATF, DNA-korjaus, DNA-vauriovaste, Fanconin anemia, ituradan
mutaatio, MRG15, PALB2, perinnöllinen rintasyöpäalttius, rintasyöpä
8
Acknowledgements
This study was carried out at the Laboratory of Cancer Genetics, Department of
Clinical Genetics, Biocenter Oulu, University of Oulu and Oulu University
Hospital during 2008–2014. I wish to express my sincere gratitude to all those
who participated in this project.
My supervisor Professor Robert Winqvist is warmly acknowledged for giving
me the opportunity to take my first steps in this fascinating scientific field and to
learn so much about cancer research. As a member of his research group, I have
had an excellent environment in which to do research. I greatly value his vast
knowledge in the field. I am much obliged to my second supervisor Docent Katri
Pylkäs; without her expertise and encouragement, I would not have completed
this thesis. I especially appreciate her insights, honesty, and readiness to help.
Docent Jukka Moilanen, Head of the Department of Clinical Genetics, and
Professor Juha Risteli, Head of the Department of Clinical Chemistry, are
acknowledged for their supportive attitude towards this PhD study.
I also owe my sincere thanks to Jaakko Ignatius and Arja Jukkola-Vuorinen
for being the most encouraging follow-up group ever. Every time we met, they
had excellent comments and ideas for the future.
The official referees Professor Helena Kääriäinen and Docent Pia Vahteristo
are thanked for reviewing this thesis and for insightful suggestions for its
improvement. My author-editor Carol Ann Pelli is acknowledged for precise
language revision of the manuscript. MA Terttu Tatti is acknowledged for the
Finnish language revision.
My collaborators and co-authors Kristiina Aittomäki, Kristiina Avela,
William Foulkes, Helmut Hanenberg, Jaana Hartikainen, A Kallioniemi, Hariet
von Koskull, Veli-Matti Kosma, Marketta Kähkönen, Satu-Leena Laasanen, Arto
Mannermaa, Katriina Parto, Carly Pouchet, Mervi Reiman, Thomas Rio Frio,
Johanna Schleutker and Mark Tischkowitz are sincerely thanked for their valuable
contributions to the research projects included in this thesis.
I am ever so grateful to all my present and former group members for a
supportive and helpful work atmosphere, for collaboration on various research
projects and for excellent company: Muthiah Bose, Raman Devarajan, SannaMaria Karppinen, Tuomo Mantere, Jenni Nikkilä, Sini Nurmenniemi, Hellevi
Peltoketo, Hannele Räsänen, Szilvia Solyom, Anna Tervasmäki, Mikko Vuorela.
Furthermore, Meeri Otsukka, Helmi Konola, Leena Keskitalo and Annika
Väntänen are thanked for fantastic technical assistance. Members of the
9
Laboratory of Clinical Genetics and the Department of Clinical Genetics are
acknowledged for their assistance. I also thank other scientists and staff from the
Kieppi building. This has been a progressive and open-minded place to work.
I thank from the bottom of my heart all of my friends for cheering me on
during the diffucult times. Their support and company over the years have been
very important to me. Erja, Minna, Marja, and Heini; finally, here I am! I thank
my dear friends Pauliina, Anna, Tiina, Elisa, Satu, Kisse, Irene, Erika, and many
more for sharing with me a rewarding life beyond the field of medicine.
To my large family, my aunts Lotta, Minna and Ulla, my mother-in-law Päivi,
mummi, my sisters Noora and Tilli and brothers Kassu and Mikael and their
wonderful spouses Antti, Toni, Sanna, and Elena, I express my deepest gratitude.
You have supported me in my weakest moments. I am especially grateful to my
parents Anna-Maija and Mikko for encouraging me to reach for the stars. I have
always felt loved.
Above all, my heartfelt thanks are owed to my extraordinary and wonderful
husband Hannu for his unfailing support, help, and love. You, together with our
three miracles Hugo, Mauno and Helka, are the most important things in my life.
I love you!
All patients and their family members who volunteered to participate in this
study are warmly thanked; without you, these projects would not have been
possible.
Financial support from the University of Oulu, Oulu University Hospital, the
National Graduate School of Clinical Investigation, the Cancer Foundation of
Northern Finland, the Finnish Cancer Foundation, the Orion-Farmos Research
Foundation, the Ida Montini Foundation, the Maud Kuistila Memorial Foundation,
and the Emil Aaltonen Foundation is gratefully acknowledged.
Turku, April 2014
10
Maria Haanpää
Abbreviations
aa
AATF/CHE1
AML
AT
ATM
ATR
amino acid(s)
apoptosis antagonizing transcription factor gene
acute myelogenous leukaemia
ataxia telangiectasia
ataxia telangiectasia mutated gene/protein
ataxia telangiectasia mutated and Rad3-related
gene/protein
BRCA1/BRCA1
breast cancer gene/protein 1
BRCA2/FAND1
breast cancer gene 2/gene for Fanconi anaemia
subtype D1
BRIP1/BACH1/FANCJ BRCA1-interacting protein C-terminal helicase 1/
Fanconi anaemia subtype J
BRCT
BRCA1 carboxy-terminal domain
CHK1/CHEK1
checkpoint kinase gene/protein 1
CHK2/CHEK2
checkpoint kinase gene/protein 2
chr
chromosome
CI
confidence interval
CSGE
conformation-sensitive gel electrophoresis
C-terminal
carboxy-terminal
DCIS
ductal carcinoma in situ
DDR
DNA damage response
DSB
double-strand break
ER
oestrogen receptor
ESE
exonic splicing enhancer
FA
Fanconi anaemia
fs
frame shift
GWAS
genome-wide association study
HER2/ERBB2
human epiderma growth factor reseptor 2
HNPCC
hereditary non-polyposis colorectal cancer
HR
homologous recombination
ICL
interstrand crosslink
IR
ionizing radiation
kb
kilobase
LOH
loss of heterozygosity
11
MLH1, -3
MLPA
MMC
MMR
MRE11
MRG15/MORF4L1
MSH2, -6
MSI
NBS
NBS1/NBS1
N-terminal
OR
p53/TP53
PALB2
PARP
PCR
PTEN
SNP
SSB
12
gene for DNA mismatch repair protein mutL homologue
1 and 3
multiplex ligation-dependant probe amplification
mitomycin c
mismatch repair
gene for meiotic recombination 11 homologue
mortality factor 4-like gene/protein 1
gene for DNA mismatch repair protein mutS
homologue 2 and 6
microsatellite instability
Nijmegen breakage syndrome
Nijmegen breakage syndrome gene (nibrin) /protein
amino-terminal
odds ratio
tumour suppressor gene/protein 53
partner and localizer of BRCA2 (FANCN) gene /protein
poly (ADP-ribose) polymerase
polymerase chain reaction
phosphatase and tensin homologue gene/protein
single-nucleotide polymorphism
singe-strand break
List of original publications
This thesis is based on the following articles, which are referred to in the text by
their Roman numerals:
I
Haanpää M, Reinman M, Nikkilä J, Erkko H, Pylkäs K & Winqvist R (2009)
Mutation analysis of the AATF gene in breast cancer families. BMC Cancer 9(1): 457.
II Rio Frio T*, Haanpää M*, Pouchet C, Pylkäs K, Vuorela M, Tischkowitz M,
Winqvist R & Foulkes WD (2010) Mutation analysis of the gene encoding the
PALB2-binding protein MRG15 in BRCA1/2-negative breast cancer families. J Hum
Genet 55(12): 842–843.
III Haanpää M, Pylkäs K, Moilanen JS & Winqvist R (2013) Evaluation of the need for
routine clinical testing of PALB2 c.1592delT mutation in BRCA negative Northern
Finnish breast cancer families. BMC Med Genet 14(1): 82.
IV Mantere T*, Haanpää M*, Hanenberg H, Schleutker J, Kallioniemi A, Kähkönen M,
Parto K, Avela K, Aittomäki K, von Koskull H, Hartikainen J, Kosma VM, Laasanen
SL, Mannermaa A, Pylkäs K & Winqvist R (2014) Finnish Fanconi anemia mutations
and hereditary predisposition to breast and prostate cancer. Manuscript.
*Shared first authorship.
In addition, some unpublished material is presented.
13
14
Contents
Abstract
Tiivistelmä
Acknowledgements
9 Abbreviations
11 List of original publications
13 Contents
15 1 Introduction
17 2 Review of the literature
19 2.1 Cancer ..................................................................................................... 19 2.1.1 Genes in cancer development ....................................................... 21 2.2 General features of breast cancer ............................................................ 23 2.2.1 Epidemiology ............................................................................... 23 2.2.2 Clinical and molecular features .................................................... 24 2.3 Inherited predisposition to breast cancer................................................. 26 2.4 High-risk susceptibility genes ................................................................. 28 2.4.1 BRCA1 ......................................................................................... 29 2.4.2 BRCA2 ......................................................................................... 31 2.5 Other genes associated with increased risk of breast cancer ................... 33 2.5.1 High-risk cancer syndromes ......................................................... 33 2.5.2 Moderate-penetrance susceptibility genes .................................... 35 2.5.3 Breast cancer predisposition genes of uncertain risk .................... 40 2.5.4 Low-penetrance breast cancer predisposition alleles ................... 41 2.6 Susceptibility genes assessed in Finnish clinical genetics services......... 43 2.7 Searching for novel breast cancer susceptibility genes ........................... 43 2.8 Molecular pathways to maintain cellular integrity .................................. 46 2.8.1 DNA damage response pathway ................................................... 46 2.9 Fanconi anaemia ..................................................................................... 47 2.9.1 DNA crosslink repair by the Fanconi anaemia pathway............... 48 2.9.2 FA genes; FANCA, FANCG and FANCI ..................................... 50 2.10 Candidate genes for breast cancer predisposition based on their
biological function .................................................................................. 52 2.10.1 AATF ............................................................................................ 52 2.10.2 MRG15 ......................................................................................... 53 3 Aims of the study
55 4 Materials and methods
57 15
4.1 Subjects ................................................................................................... 57 4.1.1 Study I .......................................................................................... 57 4.1.2 Study II ......................................................................................... 58 4.1.3 Study III ........................................................................................ 58 4.1.4 Study IV ....................................................................................... 59 4.2 DNA extraction ....................................................................................... 61 4.3 Mutation detection methods .................................................................... 61 4.3.1 Conformation-sensitive gel electrophoresis (I)............................. 61 4.3.2 High-resolution melt analysis (II, III, and IV) .............................. 61 4.3.3 Direct sequencing (I, II, III, IV) ................................................... 61 4.4 Mitomycin C sensitivity test and complementation group
analysis (IV) ............................................................................................ 62 4.5 Statistical and bioinformatical analysis ................................................... 62 4.6 Ethical issues ........................................................................................... 63 5 Results
65 5.1 Mutation screening of the AATF gene (I) ................................................ 65 5.2 Mutation screening of the MRG15 gene (II) ........................................... 65 5.3 Clinical significance of PALB2 c.1592delT screening in BRCAnegative high-risk breast cancer families (III) ........................................ 66 5.4 Complementation groups of Finnish Fanconi anaemia patients
and role of causative mutations in breast cancer susceptibility
(IV) .......................................................................................................... 67 5.4.1 Finnish FA mutations prevalence in cancer cohorts ..................... 69 6 Discussion
71 6.1 No mutations found in the AATF gene among Finnish breast
cancer families (I) ................................................................................... 71 6.2 MRG15 and hereditary predisposition to breast cancer (II) .................... 72 6.3 PALB2 c.1592delT mutation screening is clinically relevant in
high-risk breast cancer patients tested negative for
BRCA1/BRCA2 (III) ................................................................................ 73 6.4 Fanconi anaemia cases in Finland and breast cancer
predisposition (IV) .................................................................................. 76 6.5 Finding the missing heritability of breast cancer susceptibility .............. 78 7 Concluding remarks and future prospects
81 References
83 Original publications
115 16
1
Introduction
Cancer is one of the leading causes of morbidity and death in developed countries.
It is a complex disease caused by a combination of genetic, epigenetic, and
environmental factors. Up to every third individual develops cancer and every
fifth dies from it. Breast cancer is the most common malignancy affecting women.
Each year ~1.4 million women are diagnosed with breast cancer and around
500 000 breast cancer-related deaths are reported worldwide (American Cancer
Society). In Finland, the respective figures are approximately 4700 new breast
cancer cases and some 900 breast cancer-related deaths (Finnish Cancer Registry).
Although the great majority of malignancies occur sporadically due to
somatically acquired mutations, positive family history is one of the strongest
predisposing factors for breast cancer. Around 5–10% of all breast cancer cases
are caused by significant hereditary predisposition (Claus et al. 1996, Narod &
Foulkes 2004, Parkin 2004). In familial cancer cases, the inherited genetic defects
predispose to malignancy. Cancer predisposition can vary from weak to strong
depending on the penetrance of the gene defect. (Ponder 2001.) The main
indicators for hereditary disease predisposition are familial clustering of breast
cancer, tumour development at an early age, bilateral disease, and occurrence of
multiple primary tumours in the same individual (Barrett 2010, Marsh & Zori
2002). The major susceptibility genes involved in familial breast cancer are
BRCA1 and BRCA2 (FANCD1) (Miki et al. 1994, Wooster et al. 1995). Mutations
in these genes account for ~18% of all familial breast cancers (Lalloo & Evans
2012). These genes also predispose to ovarian cancer. A small portion of the
familial breast cancer cases are due to gene mutations causing certain rare
inherited cancer-predisposing syndromes: TP53 (mutated in Li-Fraumeni
syndrome), PTEN (Cowden syndrome), LKB1/STK11 (Peutz-Jeghers syndrome),
and CDH1 (hereditary diffuse gastric cancer-syndrome). Mutations in some other
known susceptibility genes, including ATM, BRIP1 (BACH1/FANCJ), CHEK2,
NBS1, and RAD50 as well as various low-risk factors account for a small fraction
of the inherited component of breast cancer. However, mutations in the currently
known breast cancer-predisposing genes explain only about 25–30% of the
hereditary component. (Antoniou & Easton 2006, Stratton & Rahman 2008,
Turnbull & Rahman 2008.) The search for additional candidate genes therefore
remains critical.
Most cancers contain hundreds or thousands of somatic mutations
(Alexandrov et al. 2013, Vogelstein et al. 2013, Loeb 2001). Accumulation of
17
genetic changes leads to transformation of normal cells into cancer cells
(Hanahan & Weinberg 2000, Nowell 1976, Ponder 2001). Epidemiological
studies have shown that environmental and lifestyle factors, such as smoking, diet,
and hormonal factors, increase cancer risk (Peto 2001). It has been suggested that
the remaining unexplained breast cancer predisposition be attributed to a
polygenic model together with environmental factors (Antoniou & Easton 2006,
Stratton & Rahman 2008). Improved knowledge of both the role of cancer
predisposition genes in normal tissue and malignancy and the additional hits that
occur in many cases to promote tumorigenesis is anticipated to revolutionize the
prevention and treatment of cancer by allowing therapeutic intervention at the
molecular level (Marsh & Zori 2002).
Most of the known breast cancer susceptibility genes are involved in the
DNA damage response (DDR) pathway. Therefore, novel breast cancer
predisposition genes might be revealed by investigating these genes further. Three
principal strategies for identifying breast cancer predisposition factors are
genome-wide linkage analysis, mutational screening of candidate genes, and
association studies (Turnbull & Rahman 2008). The relatively isolated Finnish
population with potentially fewer disease-predisposing genes than in populations
of more heterogeneous origin is likely to offer some advantages in the search for
additional low- and moderate-penetrance susceptibility genes.
The aim of this study was to identify further susceptibility genes by searching
for pathogenic mutations in the AATF and MRG15 genes. These genes were
chosen based on their involvement in the DNA damage response and their
interaction with known cancer susceptibility genes. The prevalence of the
previously identified cancer-related PALB2 c.1592delT mutation was evaluated
among high-risk breast cancer families identified at the Department of Clinical
Genetics, Oulu University Hospital, and who had received genetic counselling for
cancer susceptibility. The benefits of further genetic testing of these patients were
explored. Finally, Finnish Fanconi anaemia mutations and their relation to breast
cancer predisposition were investigated.
18
2
Review of the literature
2.1
Cancer
Cancer is characterized by clonal expansion of cells that, become malignant via a
series of genetic alterations (Fearon & Vogelstein 1990). These multiple somatic
mutations usually affect genes involved in cell differentiation and proliferation.
Through accumulation of genetic changes, the cells acquire properties typical of
the cancerous phenotype. Hanahan and Weinberg (2000) have proposed six
essential physiological changes, the hallmarks of cancer, which occur during the
transformation of a normal cell towards a malignant one: self-sufficiency in
signals stimulating growth, insensitivity to negative growth signals, evasion of
programmed cell death (apoptosis), unlimited replicative potential, capacity for
angiogenesis, and tissue invasion and metastasis. In addition, reprogramming of
energy metabolism and avoiding destruction by the immune system have been
suggested to represent other important qualities of malignant neoplasms. A
normal cell also needs two enabling characteristics for tumorigenesis. These are
genomic instability (generates random mutations) and an inflammatory state of
premalignant and malignant lesions (allows cancer cells to evade immunological
destruction). (Hanahan & Weinberg 2011.) These properties give cancer cells a
growth advantage over the general cell population (Figure 1).
19
Evading growth
suppressors
Avoiding
immune
destrucon
Deregulang
cellular
energecs
Resisng
cell death
THE
Sustaining
proliferave
signalling
HALLMARKS
Inducing
angiogenesis
Genome
instability
and
mutaon
OF
CANCER
Enabling
replicave
immortality
Acvang
invasion and
metastasis
Tumourpromong
inflammaon
Fig. 1. The six hallmarks of cancer appear with a white background, emerging
hallmarks with a light grey background and enabling characteristics with a dark grey
background. Modified from Hanahan & Weinberg 2000, Hanahan & Weinberg 2011.
Cell growth is tightly controlled. Most somatic cells do not have the capacity to
divide indefinitely. Based on epidemiological and in vitro experiments, common
adult epithelial cancers, such as those of the breast, colorectum and prostate, have
been estimated to require at least 5–7 mutations, while cancers of the
haematological system may require fewer mutations. (reviewed in Hahn et al.
1999, Stratton et al. 2009, Vogelstein & Kinzler 1993.) Given the multiple
defence mechanisms that protect cells from cancer development, this leads to a
very low likelihood that a single cell will accumulate the requisite number of
independent mutations needed for neoplastic transformation. Therefore, it has
been suggested that certain mutations affect genomic stability, thus increasing the
mutation rate in order for tumorigenesis to occur within the human lifespan.
(Loeb 1991.) Indeed, the mutation frequency has been estimated to be up to 1000fold higher in tumour cells than in normal cells, and the mutation rate is
particularly elevated in cancer predisposition syndromes, e.g. in patients with
Fanconi anaemia (Araten et al. 2005, Seshadri et al. 1987). Mutations may also
20
result in increased cell proliferation, which would give the mutated cell a growth
advantage relative to normal cells (Nowell 1976).
Dynamic equilibrium between DNA damage and efficient DNA repair must
be maintained; failure will result in malignant growth. Cancer cells can
destabilize their genome in at least two different and independent ways:
chromosomal instability, which means that most tumour cells have unusual
karyotypes with abnormal numbers of chromosomes and many structural
rearrangements, and less frequently, microsatellite instability (this was initially
described in association with hereditary non-polyposis colorectal cancer, HNPCC)
(Aaltonen et al. 1993, Read & Donnai 2001). The latter tumours are
chromosomally normal, but DNA testing shows that the normal checks on the
accuracy of DNA replication have been compromised (Gervaz et al. 2002). For
instance, 21% of ovarian cancer patients with BRCA1 mutations have been
reported to exhibit high microsatellite instability, whereas the corresponding
proportion for patients with BRCA2 mutations is 6% (Segev et al. 2013). The
changes from normal cell to cancer cell include both inactivation of tumour
suppressor genes and activation of oncogenes, giving cells a growth advantage
(Read & Donnai 2001).
2.1.1 Genes in cancer development
Cancer genes can be divided into three categories: oncogenes, tumour suppressor
genes and genome stability maintenance (also called caretaker) genes. Each
somatic mutation in the cancer cell genome can be classified as either a driver or
passenger mutation. Driver mutations are causally implicated in the process,
whereas passenger mutations are incidental, irrelevant, and do not provide any
particular clonal growth advantage. Mutations in oncogenes or tumour suppressor
genes are mostly driver mutations. (Alexandrov et al. 2013, Stratton et al. 2009.)
Oncogenes are mutated forms of normal cellular components, protooncogenes. Under normal conditions, proto-oncogenes stimulate cell growth and
differentiation. Mutations in proto-oncogenes promote their function, providing
cells with a growth advantage. Oncogenes were initially discovered in viral
genomes, as they were able to transduce normal cellular genes to become
constitutively active, thus inducing tumorigenesis. (Bishop 1981, Stehelin et al.
1976.) The protein products of proto-oncogenes are evolutionarily highly
conserved. In human cancer, proto-oncogenes are activated through genetic
alterations such as gain-of-function mutations (e.g. point mutations), gene
21
amplifications, and chromosomal rearrangements (Bishop 1991). At the cellular
level, oncogenes act dominantly, meaning that mutation in one allele alone can
enhance uncontrolled cell proliferation. Inherited mutations in oncogenes are,
however, relatively rare, presumably because such mutations remove normal
control over cell growth, thereby being lethal during embryogenesis (Frank 2001).
Tumour suppressor genes are negative regulators of cell growth (Marshall
1991, Weinberg 1991). They regulate cell proliferation either by controlling cell
division or by promoting programmed cell death. The loss-of-function mutations
in tumour suppressor genes promote malignancy. The leading evidence for the
presence of tumour suppressor genes emerged via studies on the childhood eye
tumour retinoblastoma. Knudson (1971) suggested that two mutational events
were required for retinoblastoma development. In hereditary cases, one mutation
was inherited via the germline of an individual and the second one occurred at the
somatic level. In the non-hereditary forms of the disease, by contrast, the two
events were both somatic, and thus, less likely to occur. Tumour suppressor genes
act recessively, i.e. both alleles need to be mutated. The second hit often involves
a gross chromosomal mechanism, resulting in hemi- or homozygosity of the
chromosomal region, which then leads to loss of heterozygosity (LOH). Such
chromosomal events include deletion, gene conversion, non-disjunctional
chromosome loss, and mitotic recombination. LOH can be detected by comparing
the genotypes of the tumour and the corresponding normal tissue at a given
polymorphic marker (Devilee et al. 2001). Some other possibilities for silencing
of tumour suppressor genes are deletion of both alleles (i.e. homozygous deletion)
or via epigenetic mechanisms such as transcriptional silencing through promoter
hypermethylation. However, mutations in tumour suppressor genes are not always
completely recessive since some alternate mechanisms exist. For instance,
haploinsufficiency and dominant-negative mutations might require only a single
hit. Haploinsufficiency occurs when one allele is insufficient to sustain the full
functionality produced by two wild-type alleles (Wilkie 1994) Dominant-negative
mutations alter the gene product in such way that it interferes with the function of
the wild-type protein, inhibiting its cellular effect (Payne & Kemp 2005).
Due to the growing number of tumour suppressor genes and knowledge about
the functional roles of the gene products, these genes have been further defined
and classified. Genes harbouring loss-of-function mutations can be classified as
gatekeepers or caretakers depending on their putative role in the cellular
processes (Kinzler & Vogelstein 1997). The latter are also known as genomic
stability maintenance genes. Gatekeepers are thought to directly regulate tumour
22
growth by promoting cell death or inhibiting proliferation. In the multistep
process of tumorigenesis, each cell type presumably has certain gatekeepers, the
inactivation of which is rate-limiting for tumour initiation (Kinzler & Vogelstein
1997). In contrast to gatekeepers, caretakers inhibit tumour growth indirectly by
maintaining genomic integrity (Kinzler & Vogelstein 1997). Mutations in
caretaker genes increase genetic instability, leading to an elevated mutation rate in
all genes, including tumour suppressor genes and oncogenes. Therefore, cancer
development is accelerated. Caretakers include genes involved in DNA repair and
replication such as mismatch repair (MMR) genes, and BRCA1 and BRCA2 genes
(Kinzler & Vogelstein 1998). The former are responsible for recognition and
repair of incorrectly paired nucleotides during normal replication (Fishel &
Kolodner 1995) and the latter for mitotic recombination, chromosomal
segregation, and maintenance of genomic integrity by homologous
recombination-mediated DNA double-strand break repair (Farmer et al. 2005,
Scully 2000). Normal functioning of genomic stability maintenance genes keeps
the number of genetic alterations at the minimum. In addition to gatekeepers and
caretakers, there is a third category of tumour suppressors called landscapers.
Defects in the landscapers are speculated to contribute to neoplastic
transformation by generating an abnormal stromal environment. The genetic
defect is not in the transforming cell population itself but rather in adjacent
stromal cells, which promotes malignancy of associated cells due to abnormal
intercellular signalling. (Kinzler & Vogelstein 1998.)
2.2
General features of breast cancer
2.2.1 Epidemiology
Breast cancer is the most common cancer and also the predominant cause of death
from a neoplastic disease in women (Ferlay et al. 2013). The lifetime risk of
breast cancer is one in eight females (Downs-Holmes & Silverman 2011).
Worldwide, it is estimated that 1.38 million women are diagnosed with the
disease each year, constituting about 23% of all cancer cases (Globocan 2008
2013). Breast cancer incidence shows marked geographical variation; it is much
less common in Asian than Caucasian women. Nevertheless, it is the prevailing
cancer in both developed and developing countries. In Finland, the proportion of
breast cancer comprises approximately one-third (32.8%) of all female cancers
23
(Finnish Cancer Registry 2013). The annual number of new cancer cases
diagnosed is on the rise. In 2011, a total of 4865 new breast cancer diagnoses
were made. By the end of that year, some 57 000 individuals who had had breast
cancer were still alive. Every year almost 900 breast cancer-related/caused deaths
occur, i.e. the age-adjusted mortality rate is 14.6 per 100 000 person-years. The
age-adjusted mortality rate has remained virtually unchanged since the 1950s
(14–15/100 000 person-years). Thus, a relative decline in breast cancer mortality
has been observed due to improved early diagnostics, following increased
awareness and the implementation of the population-based mammography
screening programme in 1987. The widespread administrations of adjuvant
therapies have also supported the decrease in mortality rates. (Ferlay et al. 2013,
Finnish Cancer Registry 2013, Mouridsen 1993.) The screening programme is
currently directed to women aged 50–69 years and the examination is performed
every other year. Almost 90% of invited Finnish women participate in the
programme. (National Institute for Health and Welfare (THL) 2013.)
2.2.2 Clinical and molecular features
The majority of breast cancer cases (81%) occur in women aged 50 years or over.
The risk of breast cancer rises throughout a woman’s lifetime. Besides female
gender, age, and family history of the disease, significant risk determinants
include several reproductive and hormonal factors, such as early menarche, late
menopause, late age at first childbirth, nulliparity or low number of children,
short breast-feeding, use of oral contraceptives, and hormone replacement therapy,
all of which increase the exposure of breast tissue to oestrogen (Barrett 2010).
Oestrogens have an important role in the development and progression of breast
cancer. Both direct and indirect mechanisms support oestrogen’s contribution to
the initiation and promotion of breast cancer. Furthermore, some oestrogen
metabolites have been reported to have genotoxic properties. Breast tissue acting
as an intracrine organ with local oestrogen production may also play some role in
carcinogenesis, but this is poorly understood (reviewed in Yaghjyan & Colditz
2011). Other breast cancer risk factors are postmenopausal obesity, poor nutrition,
alcohol/tobacco consumption, lack of exercise, low socio-economic status,
exposure to radiation, and a previous benign breast disease (Downs-Holmes &
Silverman 2011, Folsom et al. 1990).
Breast tumorigenesis is a multi-step process in which normal breast
epithelium evolves via hyperplasia and carcinoma in situ into an invasive cancer.
24
About 95% of malignant breast tumours are carcinomas from the epithelium of
the mammary gland. (Berg & Hutter 1995.) Histologically, the most common
breast cancer tumours are infiltrating ductal (75%) and lobular (15%) carsinomas,
and less common forms include medullar (1–7%), mucinous (1–2%), tubular (1–
2%), papillary (>1%) and inflammatory carcinomas. The different histological
types have different prognoses. For example, lobular breast cancer has a slightly
better prognosis than a ductal tumour. Of the rarer subtypes, tubular carcinoma is
a well-differentiated carcinoma with an excellent prognosis, whereas rare
inflammatory carcinoma has a poor prognosis. (Downs-Holmes & Silverman
2011, Li et al. 2003, Li et al. 2005.)
The differentiation of the tumour cells defines the tumour histological grade.
Grading is performed using the method by Bloom and Richardson (1957),
subsequently modified by Elston and Ellis (1991). The histological grade is based
on three features: tubule formation, nuclear pleomorphism, and mitotic count.
Eventually, breast cancer stage becomes the most important prognostic factor,
forming the basis for selecting patients for different treatment strategies. The
staging is performed by TNM classification, which describes the size of the
tumour (T), involvement of local lymph nodes (N), and distant metastasis (M).
The routine pathological classification of a breast tumour also reveals lymphatic
or vascular channel invasion, infiltration into skin, presence and amount of
oestrogen (ER) and progesterone (PR) receptors, and expression of the HER2
(also called ERBB2 or NEU) protein. The proliferation rate of tumour cells and
the expression of the tumour suppressor protein p53 are also frequently
determined. (UICC 2013.)
Based on mRNA microarray-based gene expression patterns, breast cancers
are often divided into five molecular subtypes: luminal A, luminal B, HER2enriched, basal-like, and normal-like. These subtypes have preferential sites for
distant relapse, and the subtype may also be associated with efficacy of systemic
cancer therapies. The basal-like breast cancer type includes triple-negative breast
cancer (ER-, PR-, HER2-negative), which shows more aggressive clinical
behaviour, has the highest cell proliferation rates, and has a poor prognosis. This
subtype accounts for 15% of all breast carcinomas. Most breast cancers belong to
the luminal A (71%) and B (8%) subtypes, and these tumours are oestrogen- and
progesterone receptor-positive and are associated with a more favourable
prognosis since they have low cellular proliferation rates. (Hugh et al. 2009,
Rouzier et al. 2005, Smid et al. 2008, Sorlie et al. 2001, Wiechmann et al. 2009.)
25
Although breast cancer incidence is increasing, the prognosis has improved,
partly because of earlier diagnosis and partly as a result of the use of adjuvant
therapies. Four types of standard treatments are used for breast cancer: surgery,
radiation therapy, chemotherapy, and hormone therapy. Radiation therapy,
chemotherapy, and hormone therapy are often used in conjunction with surgery.
The five-year relative survival rate (RSR) is nowadays as high as 86% (Finnish
Cancer Registry 2013).
Breast cancer also occurs in males, albeit rarely. In 2011, a total of 17 new
male breast cancer cases were diagnosed in Finland. This comprises 0.11% of all
cancers in men (Finnish Cancer Registry 2013). Germline mutation of BRCA2 is
the greatest known risk factor for male breast cancer. An estimated 13% of
Finnish male breast cancer patients are BRCA2 mutation carriers. (Syrjäkoski et al.
2004.)
2.3
Inherited predisposition to breast cancer
In 1948, Penrose and colleagues observed that in some families breast cancer may
have a hereditary basis (Penrose et al. 1948). A genetic influence on breast cancer
susceptibility is suggested by twin studies, as the concordance for breast cancer in
identical twins (0.28) is more than twice that in dizygotic twins (0.12) (Petrakis
1977). Hereditary breast cancer is classified by a clustering of breast cancers in a
family that is substantially higher than the observed incidence in the general
population (average incidence in a standard Western population for breast cancer
is 88.4/100 000 new cases per year (Globocan 2013). Other factors linked to
hereditary breast cancer susceptibility are early disease onset, occurrence of
bilateral breast cancer or multiple primary tumours in the same individual, or
male breast cancer in the family. Familial breast cancer is usually seen in several
generations in a pedigree. (Thull & Vogel 2004.)
About 5–10% of breast cancers are caused by a hereditary predisposition to
the disease, i.e. a strong familial background of breast and/or ovarian cancer
exists (Claus et al. 1996, Narod & Foulkes 2004, Parkin 2004). Interestingly,
genetic factors and inherited germline mutations may play a role in up to 27%
(0.04–0.41 CI) of breast cancers according to data from the Nordic twin registries
(Lichtenstein et al. 2000). First-degree relatives of a breast cancer patient have an
approximately twofold risk for breast cancer compared with the general
population; thus, positive family history is considered one of the strongest
predisposing factors for breast cancer. The risk increases with the number of
26
affected relatives. (Collaborative Group on Hormonal Factors in Breast Cancer
2001, Thompson & Easton 2004.)
Breast cancer susceptibility genes have been divided into three different
classes: rare high-penetrance genes, rare intermediate/moderate-penetrance genes,
and common low-penetrance genes. This classification has been made according
to the level of breast cancer risk conferred and the prevalence of disease-causing
mutations in the general population. According to the classification proposed by
Stratton and Rahman (2008), high-penetrance genes confer a 10- to 20-fold
relative breast cancer risk and the mutation frequency in the population is very
low. Moderate-penetrance genes generally increase the risk 2- to 4-fold and the
frequency is low, and finally, low-penetrance alleles usually confer 1.5-fold risk
or less but their frequency in the general population is high. The level of
penetrance describes the likelihood that an individual who carries a certain
mutation will develop a disease due to its presence. Breast cancer susceptibility
genes have been identified by genome-wide linkage analysis and positional
cloning, mutation screening/next generation sequencing and genome-wide
association studies (GWASs) (reviewed in Turnbull & Rahman 2008). Mutations
in all of the currently identified susceptibility genes account only for about 25–30%
of familial breast cancers (Ellsworth et al. 2010, Turnbull & Rahman 2008).
Hence, much work is still needed to understand the genetic background of breast
cancer. More than 70% of high-risk families for inherited genetic predisposition
to breast cancer remain without an explanation (Figure 2).
27
Fig. 2. Affected genes in hereditary breast cancer. Modified from (Lalloo & Evans 2012,
van der Groep et al. 2011).
The polygenic theory of breast cancer susceptibility suggests that the residual
unexplained familial cancer may be due to several genetic variants. Each variant
alone may cause only a small and often undetectable increase in cancer risk, but
together their effect could be sufficient to have a significant input on cancer
development. These effects are also likely to be influenced by environmental and
epigenetic factors. (Antoniou et al. 2003, Antoniou & Easton 2006, Pharoah et al.
2002.)
2.4
High-risk susceptibility genes
Fortunately, mutations in high-penetrance genes at the population level are very
rare (<0.1%) since they are associated with a very high breast cancer risk (up to
20-fold). An estimated 10–30% of all breast cancers are attributed to hereditary
factors, but only 5–10% of the cases display a strong inherited component, and
further, less than one-quarter (22%) of these heritable cases would be explained
by mutations in high-penetrance genes transmitted in an autosomal dominant
28
manner (Apostolou & Fostira 2013, Ellsworth et al. 2010). The two most
important high-risk genes are BRCA1 and BRCA2, together explaining about 16%
of all familial breast cancers (Anonymous 2000, Peto et al. 1999). An example of
a pedigree showing an apparently autosomal dominant susceptibility to breast
cancer was first described already in 1866 by a French physician, Paul Broca,
whose wife’s family contained 10 women in four generations who died of the
disease (Broca 1866, Poumpouridou & Kroupis 2011).
2.4.1 BRCA1
By performing a genetic linkage analysis in 23 early-onset breast cancer families,
the BRCA1 gene was first localized to chromosome 17q21 (Hall et al. 1990) and
subsequently cloned (Miki et al. 1994). The pathogenic mutations in BRCA1 were
found to account for about 7–10% of all familial breast cancers (Peto et al. 1999).
BRCA1 consists of 24 coding exons and the 7.4 kb transcript encodes a protein of
1863 amino acids. BRCA1 does not have an apparent close homologue in the
human genome, not even BRCA2, although both proteins share some
characteristic structural features (Thompson & Easton 2004). Highly conserved
functional motifs of BRCA1 include some important domains: the N-terminal
RING finger domain and nuclear export signal (NES), two BRCT domains at the
C-terminus, and a coiled-coil domain preceding the BRCT domains. (Caestecker
& Van de Walle 2013, Koonin et al. 1996.) BRCA1 is associated with a series of
protein complexes, and numerous proteins are known to bind to the functional
domains of BRCA1. Its structure and the binding sites of interacting proteins are
shown in Figure 3.
The BRCA1 protein is involved in several important cellular pathways
maintaining genomic stability, including double-strand break (DSB) repair by
homologous recombination (HR), cell cycle checkpoint control, DNA replication,
transcription regulation, nucleotide excision repair, protein ubiquitylation, and
chromosome duplication, as well as histone deacetylation during chromatin
remodelling. Indeed, BRCA1 is involved in several DNA repair pathways,
including both DSB and single-strand break repair (reviewed in Wang 2012,
Caestecker & Van de Walle 2013). BRCA1 appears to have an early role in the
regulation and promotion of HR, which repairs DSBs and interstrand crosslinks
(ICL). BRCA1 is phosphorylated in response to DNA DSBs by several kinases,
e.g. ATM (ataxia telangiectasia mutated), CHK2, and/or ATR (ataxia
telangiectasia Rad3-related). DNA damage triggers the ATM- and ATR-dependent
29
phosphorylation of H2AX (γH2AX), which is required for accumulation of
numerous DNA damage response proteins into subnuclear foci at DSBs and for
ensuring optimal repair. BRCA1 forms several different types of complexes at
sites of damage. At least five different protein complexes are known. Core
complex is formed by BRCA1 and BARD1 in the RING domain, which possesses
E3 ubiquitin ligase activity. BRCA1 exists as a stable heterodimer with BARD1,
and BARD1 is present in all BRCA1 complexes. BRCA1/PALB2/BRCA2
supercomplex is formed through binding of PALB2 with the coiled-coil domain
of BRCA1, and it plays a role in homologous recombination-mediated DNA
repair. BRCA1-A contains ABRAXAS, RAP80, NBA1/MERIT40, BRE/BRCC45,
and BRCC36 proteins, and it is associated with ubiquitylated histones near the
sites of DNA damage. The BRCA1-B complex includes BRIP1 and TOPBP1 and
this complex is required for replication stress-induced checkpoint control and
DNA interstrand crosslinks repair. The BRCA1-C complex consists of CtIP and
the MRN (MRE11/RAD50/NBS1) proteins, and it senses DSBs and is responsible
for DNA end resection. All three of these complexes (A, B, C) are formed through
interactions with the BRCT domains of BRCA1. (Boulton 2006, Huen et al. 2010,
Li & Greenberg 2012, Rosen 2013, Roy et al. 2011, Tutt & Ashworth 2002, Wang
2012, Welcsh et al. 2000.)
The clinically important mutations of BRCA1 most often target the RING and
BRCT domains, and the majority of the mutations are small frameshift insertions
or deletions, resulting in prematurely truncated protein products. This means that
this kind of mutation, if able to escape nonsense-mediated decay at the mRNA
level, generates shortened, non-functional BRCA1 proteins that are far less stable
and more susceptible to proteolytic degradation. (Caestecker & Van de Walle
2013, Narod & Foulkes 2004, Thompson & Easton 2004.) There are some
founder mutations in specific populations, but generally a preponderance of
BRCA1 mutations is rare, and they are reported to occur only in single families
(~60% just once) (Thompson & Easton 2004, Turnbull & Rahman 2008).
The average cumulative risk for breast cancer in BRCA1 mutation carriers by
age 70 years has been estimated to be 65% (95% confidence interval 44–78%)
and for ovarian cancer 39% (95% CI 18–54%). The relative risk of breast cancer
peaks in the age group 30–39 years, declining thereafter, while ovarian cancer RR
estimates show no apparent trend with age. On the other hand, breast cancer
incidence in BRCA1 mutation carriers increases with age up to 45–49 years,
remaining roughly constant thereafter. (Antoniou et al. 2003.) Bilateral cancer is
common in BRCA1 mutation carriers. The estimated population frequency of
30
mutations in this gene is about 1/1000 per gene in the United Kingdom (Turnbull
& Rahman 2008).
BRCA1-associated breast cancers seem to be of a higher grade, displaying a
high degree of pleomorphism and an elevated mitotic frequency. The tumours are
often ER-, PR-, and HER2-negative, p53-positive, aneuploid, and of medullary
histology. Tumours in BRCA1 carriers share a similar immunohistochemical
profile to sporadic basal carcinomas. The data from different studies regarding
prognosis of BRCA1-associated breast cancers are confusing, suggesting from
better to worse outcome compared with sporadic cancers, although the association
of basal cancer with poorer prognosis would support the hypothesis of worse
prognosis for patients with BRCA1 mutation. (reviewed in Hodgson 2006,
Lakhani et al. 2005, Lalloo & Evans 2012, Mavaddat et al. 2012.)
Fig. 3. Schematic structure and interacting proteins of BRCA1 and BRCA2 (more
details in the text). The interacting proteins are shown below the BRCA region
required for their association (Roy et al. 2011). Presented with permission from the
Nature Publishing Group.
2.4.2 BRCA2
The BRCA2 gene was first chromosomally localized by analysing 15 high-risk
breast cancer families that did not show linkage to the BRCA1 locus. This
31
genome-wide linkage analysis mapped BRCA2 to chromosome 13q12-q13 in
1994, and the gene was cloned the following year (Wooster et al. 1994, Wooster
et al. 1995). Pathogenic mutations in BRCA2 account for approximately 10% of
all familial breast cancer cases (Anonymous 2000, Lalloo et al. 2003).
BRCA2 is a large gene with 27 exons and spans about 70 kb of genomic DNA
and encodes a protein of 3418 amino acids. BRCA2 contains eight BRC repeats
between amino acid residues 1009 and 2083, and one DNA-binding domain,
which contains a helical domain, three oligonucleotide binding folds, and a tower
domain at the C-terminus (reviewed in Roy et al. 2011). The structure of BRCA2,
showing also the binding sites of its interacting proteins, is presented in Figure 3.
In contrast to the more peripheral roles of BRCA1 in regulation of the core
HR machinery, BRCA2 plays a more central role in DNA repair via regulation of
the activity of the RAD51 (Turner et al. 2005). BRCA2 has very important
functions in DDR since it facilitates HR and is involved in DSB repair. BRCA2
binds to RAD51 through eight evolutionarily conserved RAD51 binding domains,
termed the BCR repeats, and a separate motif located at its C-terminal domain.
BRCA2 regulates the formation of RAD51 nucleoprotein filaments (Davies et al.
2001, Pellegrini et al. 2002, Sharan et al. 1997). RAD51 is a key protein that
coats the processed single-stranded DNA overhangs of DSBs and promotes
homologous pairing and strand invasion of these regions during homologous
recombination (West 2003). BRCA2 can also bind to single-stranded DNA via a
C-terminal domain, the structure of which is critical to the ability of BRCA2 to
promote recombination. Following DNA damage and initial DSB processing,
BRCA2 re-localizes to the site of DNA damage (Yu et al. 2003, Yuan et al. 1999).
BRCA2 acts preferentially at the interface between double-stranded DNA and the
MRN-processed single-stranded 3’ overhang to displace replication protein from
the overhang and facilitate the loading of RAD51 (Yang et al. 2002, Yang et al.
2005). The RAD51 nucleoprotein filament subsequently catalyses the search for
identical target sequences and strand invasion. Binding of RAD51 to BRCA2 is
regulated by cell cycle (CDK)-dependent phosphorylation and appears to function
as a “switch” controlling recombinational repair activity during the transition
from S/G2 to M phase in the cell cycle (Esashi et al. 2005).
The average cumulative risk of breast cancer for BRCA2 mutation carriers by
age 70 years is 45% (95% CI 31–56%), and for ovarian cancer 11% (2.4–19%).
The estimated relative risk of breast cancer among mutation carriers is highest in
the youngest age group (20–29 years) thereafter decreasing, while the ovarian
32
cancer RR starts to increase from age 40 until age 59, then decreasing (Antoniou
et al. 2003).
In addition to breast and ovarian cancer predisposition, BRCA2 is also a
Fanconi anaemia (FA) gene. Biallelic mutations in BRCA2 result in FA subtype
D1 (OMIM 605724) (Howlett et al. 2002). (see Sections 2.7 and 2.8.2 for more
detailed discussion). In contrast, biallelic BRCA1 mutations have been reported
convincingly in humans only once, where a woman was diagnosed with earlyonset ovarian cancer and had validated deleterious biallelic mutations in the
BRCA1 gene (Domchek et al. 2013). Earlier, it was presumed that biallelic
BRCA1 mutations are embryonically lethal (Evers & Jonkers 2006).
2.5
Other genes associated with increased risk of breast cancer
2.5.1 High-risk cancer syndromes
Breast cancer is also a component of several cancer syndromes, including LiFraumeni syndrome (OMIM 151623), Cowden syndrome (OMIM158350), and
Peutz-Jeghers syndrome (OMIM 175200). These three syndromes are rare, with
carrier frequencies <0.1% (Turnbull & Rahman 2008). They explain only around
~2% of the high-risk breast cancer cases, and these high-penetrance genes are
associated with a breast cancer relative risk higher than 5 (Bradbury & Olopade
2007, Stratton & Rahman 2008, Thompson & Easton 2004).
The p53 tumour suppressor gene, i.e. TP53, is mutated in the Li-Fraumeni
syndrome (Malkin et al. 1990, Srivastava et al. 1990), which is characterized by
the occurrence of multiple cancers, including childhood soft tissue sarcomas,
osteosarcomas, leukaemia, brain tumours, adrenocortical carcinomas, and earlyonset breast cancer. Mutations in TP53 are found in approximately 70% of the LiFraumeni families and predisposition to this syndrome is inherited in a autosomal
dominant manner (Malkin et al. 1990, Srivastava et al. 1990). Based on follow-up
studies, the lifetime malignancy risk in female mutation carriers is higher and age
at onset is lower than in male carriers (Kast et al. 2012, Ross & Hortobagyi 2005).
On average, 30% of the female mutation carriers develop breast cancer by the age
of 30 years (Malkin et al. 1990, Srivastava et al. 1990) and 50–60% are affected
by the age of 45 years (Chompret et al. 2000). The lifetime penetrance of TP53
mutations for cancer approaches 100% (Bradbury & Olopade 2007). However,
Li-Fraumeni syndrome accounts for less than 0.1% of all breast cancer cases
33
(Lalloo & Evans 2012). Although only a few germline mutations in TP53
contribute to hereditary breast cancer, somatic inactivation of TP53 in tumours is
frequently identified. It is among the most frequently mutated genes in human
cancer (Sengupta & Harris 2005). The TP53 gene is generally called the guardian
of the genome. Inactivation of p53 prevents the initiation of apoptosis, thus
contributing to the establishment and progression of malignancy (angiogenesis
and invasion) (Muller & Vousden 2013). Knowledge of the TP53 mutational
status could be important when predicting tumour response to radiotherapy
(Mirzayans et al. 2012). Gene-profiling of tumour tissue for more precise
selection of cancer treatment is already used in a private clinic in Finland
(Docrates syöpäsairaala).
Multiple hamartomatous lesions in a variety of tissues from all three
embryonic layers, especially of the skin, mucous membranes, breast, thyroid, and
endometrium are characteristic of Cowden syndrome. The syndrome has been
described as early as in 1963 by Lloyd and Dennis (1963) and then PTEN
(phosphatase and tensin homologue on chromosome 10) mutations were
identified as causative in 1996 (Nelen et al. 1996). It is an autosomal dominant
cancer susceptibility syndrome, and mutations of PTEN are found in over 80% of
Cowden disease families (Mallory 1995, Nelen et al. 1996, Ross & Hortobagyi
2005). Women with Cowden syndrome have a 25–50% lifetime risk of
developing malignant breast disease, with an average age of diagnosis between 38
and 46 years (Brownstein et al. 1978, Starink et al. 1986). Similarly to the TP53,
germline mutations in PTEN are rare, but PTEN is frequently inactivated in
sporadic cancers. The three most commonly sporadic cancers harbouring
mutations in the PTEN gene are glioblastoma, prostate cancer and endometrial
cancer (Chow & Baker 2006). PTEN is involved in the regulation of apoptosis
and angiogenesis (Hobert & Eng 2009, Ross & Hortobagyi 2005).
A serine-theorine kinase encoding the STK11/LKB1 gene (located at 19p13.3)
has been found to be mutated in several families with Peutz-Jeghers syndrome
(Hemminki et al. 1998). This is an autosomal dominant disorder characterized by
multiple gastrointestinal hamartomatous polyps, mucocutaneous pigmentation
(especially on the border of lips), and an increased risk for various neoplasms,
including breast and gastrointestinal cancer. The cumulative risk for breast cancer
was estimated to be 45% (27–68%) by the age of 70 years (Hearle et al. 2006).
The function of STK11 is complex. It plays a role in cellular energy metabolism,
regulates cellular proliferation, and influences cell polarity, cell growth and p53
mediated apoptosis (reviewed in Beggs et al. 2010).
34
Hereditary diffuse gastric cancer is caused by germline mutations in cell-tocell adhesion protein E-cadherin encoded by gene (CDH1), which is located at
chromosome 16q22.1. (Guilford et al. 1998). Mutation carriers have a more than
70% lifetime risk of developing diffuse gastric cancer and a 39–52% risk of
lobular breast cancer (Guilford et al. 2010, Kaurah et al. 2007, Pharoah et al.
2001). In cancer development, the most common mechanism of downregulation is
promoter hypermethylation of the second copy of the CDH1 gene (Guilford et al.
2010). Loss of E-cadherin affects the maintenance of cell architecture and results
in an abnormal orientation of the mitotic spindle (Humar & Guilford 2009).
2.5.2 Moderate-penetrance susceptibility genes
Moderate-risk variants have been estimated to account for about 5% of excessive
familial breast cancer risk. Carriers of moderate- (also called intermediate-)
penetrance alleles have a two- to four-fold risk of developing breast cancer
compared with the general population riks of 10% (Hollestelle et al. 2010,
Stratton & Rahman 2008). The prevalence of the variants is usually rare, i.e. they
have ~1% frequency in the general population, but it varies widely among
different geographical or ethnic populations. In contrast to high-risk genes,
moderate-risk breast cancer alleles typically have a limited number of variants
that confer breast cancer risk. Inheritance of moderate-risk alleles will not cause a
highly penetrant disease pattern among families, and therefore, inheritability is
not necessarily easy to recognize. Moderate-risk breast cancer genes identified
thus far have been discovered through candidate gene association studies,
involving evaluation of putative breast cancer genes in cohorts of hundreds of
cases and controls. (Hollestelle et al. 2010.)
To date, six moderate-risk breast cancer genes have been identified: CHEK2,
ATM, BRIP1, NBS1, RAD50, and PALB2. On average, mutations in these genes
confer a relative cancer risk between 1.5 and 5, the typical increase in risk being
2- to 3-fold. However, some of them might also harbor mutations that confer
higher increased risks, e.g. PALB2 which mainly confers a 2- to 6-fold increased
risk, or even higher (Byrnes et al. 2008, Shuen & Foulkes 2011, Southey et al.
2010, Turnbull & Rahman 2008, Walsh & King 2007).
The CHEK2 (checkpoint kinase 2) gene is located on chromosome 22q12.1
and encodes a cell cycle checkpoint kinase, which is a key mediator in the DNA
damage response. This kinase is responsible for arresting mitosis in response to
DNA damage by phosphorylating a number of proteins involved in cellular
35
checkpoint control, including BRCA1, p53, and CDC25C. (Matsuoka et al. 1998,
Shuen & Foulkes 2011, Wu et al. 2001.) A founder mutation, 1100delC, which
abrogates the kinase domain, has been discovered in Northern European
populations (Meijers-Heijboer et al. 2002, Vahteristo et al. 2002). This mutation
is associated with a 2.34-fold risk of breast cancer (CHEK2 Breast Cancer CaseControl Consortium 2004). However, a meta-analysis has shown that a carrier of
a CHEK2 mutation with a strong family history of breast cancer is at comparable
risk of breast cancer to BRCA carriers: an estimated lifetime risk of 37%
(Weischer et al. 2008).
The ATM (ataxia telangiectasia mutated) gene is located on chromosome
11q22.3 and encodes a checkpoint kinase that plays a role in DNA DSB sensing
and signalling. ATM also regulates several tumour suppressors, including p53,
BRCA1, and CHEK2. Biallelic (homozygous or compound heterozygous)
mutations in this gene are linked to a rare human autosomal recessive disorder
called ataxia telangiectasia (AT), which is characterized by progressive cerebellar
ataxia, a weakened immune system, hypersensitivity to ionizing radiation, highly
increased susceptibility to malignancies, primarily of lymphoid origin, and
distinctive dilated blood vessels in the eyes and skin (telangiectasia). (Ahmed &
Rahman 2006, Easton 1994, Savitsky et al. 1995, Swift et al. 1976, Thompson et
al. 2005.) A heterozygous mutation of ATM does not lead to the AT phenotype,
but carriers have an approximately 2.37 relative risk of breast cancer (Renwick et
al. 2006, Thompson et al. 2005). Markedly higher risks, up to 15-fold, have also
been reported (Chenevix-Trench et al. 2002, Pylkäs et al. 2007). There are many
ATM variants that confer a breast cancer risk, including not only all truncating
variants but also a variety of missense variants (reviewed in Hollestelle et al.
2010). In the Finnish population germline mutations in ATM have an apparent but
limited contribution to familial and sporadic breast cancer predisposition outside
AT families (Pylkäs et al. 2007). The gene mutation’s penetrance is approximately
15%, but accurate prediction of which mutation carriers will develop breast
cancer is not feasible (Apostolou & Fostira 2013), similar to all other currently
known predisposing genes.
The BRIP1/BACH/FANCJ gene (BRCA1-interacting protein C-terminal
helicase 1) is located in chromosome 17q22-q24 and encodes a DEAH helicase
that interacts directly with BRCA1 C-terminal BRCT repeats and has BRCA1dependent roles in DNA repair and checkpoint control (Cantor et al. 2001, Peng
et al. 2006). In 2006, truncating BRIP1 mutations were identified in breast cancer
families (Seal et al. 2006). Heterozygous female carriers of truncating BRIP1
36
mutations have an approximately two-fold increased breast cancer risk, with
higher risks for women younger than 50 years. As expected for a moderate-risk
gene, co-segregation of the mutation with the breast cancer phenotype is
incomplete (Hollestelle et al. 2010, Seal et al. 2006). BRIP1 germline mutations
also confer an increased risk of ovarian cancer (Rafnar et al. 2011). Concurrently,
it emerged that biallelic mutations in BRIP1 result in Fanconi anaemia
complementation group J (OMIM 609054). This phenotype is different from that
caused by biallelic mutations in BRCA2, resulting in a much lower rate of
childhood solid tumours (Levitus et al. 2005, Levran et al. 2005, Litman et al.
2005).
The MRE11-RAD50-NBS1 (MRN) protein complex plays an important role
in primary sensing and early processing of DSBs by promoting ATM’s
localization to DSB sites to initiate the DNA damage response. This protein
complex integrates DNA repair with checkpoint signalling through ATM, BRCA1,
and CHEK2 proteins. (Bartkova et al. 2008, Lavin 2007, reviewed in DzikiewiczKrawczyk 2008). Biallelic hypomorphic germline mutations of NBS1 manifest as
Nijmegen breakage syndrome (NBS) (OMIM 251260). NBS is a rare autosomal
recessive disorder characterized by microcephaly, immunodeficiency, and
increased predisposition to malignancies, especially leukaemia and lymphoma at
a young age. Relatives of NBS patients carrying heterozygous mutations are also
more prone to cancer development, including breast cancer. (Seemanova 1990,
Seemanova et al. 2007.) Mutations in MRE11 lead to an ataxia-telangiectasia-like
disorder (OMIM 604391), which is a very rare autosomal recessive disease with
ocular apraxia and ataxia in infancy (Palmeri et al. 2013). In contrast to NBS1 and
MRE11 mutations, so far only one patient has been reported with biallelic
hypomorphic mutations in RAD50. This patient exhibited phenotypes similar to
NBS (Lamarche et al. 2010). All three MRN complex genes are suggested to have
a role in breast cancer susceptibility. NBS1 variants have been found to be
associated with an increased risk of developing cancer; the Slavic founder
mutation NBS1 657del5 is particularly implicated in predisposition to breast
cancer. It is associated with a 3.1-fold enrichment among breast cancer cases
relative to those without the allele (Bogdanova et al. 2008, di Masi & Antoccia
2008, Steffen et al. 2006). For RAD50, the Finnish 687delT founder mutation has
been reported to be associated with a 4.3-fold increased breast cancer risk in allele
carriers (Heikkinen et al. 2003, Heikkinen et al. 2006). By contrast, the role of
MRE11 in breast cancer remains unclear. The potential clinical relevance of the
MRE11 and the whole MRN complex in human breast cancer is highlighted by
37
the observation that a subset of basal-like human breast cancers has markedly
reduced expression/loss of MRN-complex proteins. Subsequent sequencing of the
MRN genes from the tumours of altogether eight patients revealed two germline
mutations in the MRE11 gene (Bartkova et al. 2008). However, more studies are
needed to determine the role of MRE11 as a breast cancer susceptibility gene.
PALB2
PALB2 (partner and localizer of BRCA2) was first identified through its
interaction with BRCA2. The PALB2 gene is located in chromosome 16p12 and it
consists of 13 exons, which encodes a protein of 1186 amino acids (Xia et al.
2006). The PALB2 structure, starting from the N-terminus, contains a coiled-coil
domain, which directly binds to BRCA1 (aa 9-44), followed by two DNA binding
sites (P2T1 and P2T3), and a ChAm (chromatin-association) binding motif.
Seven WD40-repeats interact with BRCA2 (aa 853-1186), whereas the interaction
with RAD51 is mediated through two different molecular sites: aa 1-200 and aa
836-1186. (Bleuyard et al. 2012, Buisson et al. 2010, Dray et al. 2010, Livingston
2009, Oliver et al. 2009, Sy et al. 2009b, Sy et al. 2009c, Zhang et al. 2009a,
Zhang et al. 2009b.) PALB2 participates in the regulation of the nuclear functions
of BRCA2. Particularly, it promotes the stable intranuclear localization or
accumulation of BRCA2, which facilitates BRCA2 functions in homologous
recombination or DSB repair and the S phase checkpoint. BRCA2 localization to
the damaged sites actually occurs via a PALB2-BRCA1 interaction; thereby
PALB2 physically links BRCA1 and BRCA2 proteins to form a “BRCA”
supercomplex. BRCA1 has been shown to be an upstream regulator of both
PALB2 and BRCA2 in DDR, as BRCA1 promotes their concentration at the sites
of DNA damage. Overall, the association between BRCA1 and PALB2 is
important for HR-mediated DBS repair, and PALB2 is essential for key nuclear
caretaker functions of BRCA2, since it ensures proper BRCA2 nuclear presence.
(Sy et al. 2009b, Xia et al. 2006, Zhang et al. 2009a, Zhang et al. 2009b.)
Different studies have identified MRG15 as a component of certain chromatin
remodelling complexes that bind to PALB2 and have also suggested that this
interaction is required for recruiting the entire BRCA supercomplex to the
damage sites (Hayakawa et al. 2010, Sy et al. 2009a).
Germline mutations in PALB2 are associated with an increased risk of breast
cancer (Erkko et al. 2007, Rahman et al. 2007). Mutations have been identified in
approximately 1% of hereditary breast cancer families throughout the world.
38
However, the frequencies vary in different populations, ranging from 0.1% to
2.7%, and the average penetrance of deleterious PALB2 mutations is not known
with certainty (Tischkowitz & Xia 2010). One of the first two studies showing an
association between germline mutations in PALB2 and breast cancer estimated the
relative risk to be 2.3. Interestingly, the risk in carriers appeared relatively higher
in patients diagnosed at a younger age, but this requires confirmation with
additional studies. (Rahman et al. 2007.) A simultaneously published Finnish
study suggested that the risk conferred by PALB2 mutations might be significantly
higher (Erkko et al. 2007). In support of the Finnish study, a study from Australia
reported the hazard ratio of the PALB2 c.3113G>A mutation to be as high as 91%
(95% CI 44–100%) by age of 70 years (Southey et al. 2010). Most pathogenic
PALB2 mutations detected so far are truncating frameshift or stop codons and are
scattered throughout the entire gene region with no hot-spot areas (Poumpouridou
& Kroupis 2011). Except for the founder mutations, PALB2 alterations appear to
be individually rare and may confer only a small fraction of the familial risk of
breast cancer. PALB2 mutations have also been associated with familial pancreatic
cancer (Jones et al. 2009).
Analogous to BRCA2, PALB2 is also a Fanconi anaemia gene. Soon after
PALB2 was discovered, biallelic mutations were identified in eight FA families,
which identified PALB2 as a FA gene, named FANCN, causing FA subtype N (FAN). The phenotypes of FA-N and that caused by BRCA2 mutations (FA-D1) are
very similar and can be distinguished from classical FA by markedly elevated
frequency of early childhood solid tumours such as Wilms’ tumour and
medulloblastoma. (Reid et al. 2007, Xia et al. 2007.) The similarity of phenotypes
of FA-D1 and FA-N individuals could reflect the close functional relationship that
exists between BRCA2 and PALB2, both having vital roles in HR repair, with
dysfunction in either one of these proteins leading to fundamental defects in DSB
repair, a feature that is not consistent with other FA subtypes. (Reid et al. 2005,
Reid et al. 2007.)
PALB2 1592delT mutation
Several studies worldwide have demonstrated that monoallelic truncating PALB2
mutations are associated with breast cancer susceptibility (Cao et al. 2009,
Foulkes et al. 2007, Garcia et al. 2009b, Rahman et al. 2007, Southey et al. 2010).
In Finland, one recurrent PALB2 mutation has been reported (Erkko et al. 2007).
The prevalence of this PALB2 c.1592delT founder mutation has been
39
approximated to be 1% in unselected breast cancer cases and 2.6% in those with a
strong history of breast cancer in the family (Erkko et al. 2008, Heikkinen et al.
2009). The key function of PALB2 is to interact with BRCA2 and to facilitate
appropriate communication between BRCA1 and BRCA2 (Hamel et al. 2008,
Zhang et al. 2009b), which is apparently disturbed by c.1592delT (Nikkilä et al.
2013). PALB2 c.1592delT is a loss-of-function mutation that, results in a
truncated protein product with reduced stability, and consequently, the full-length
PALB2 protein levels are significantly decreased in mutation carrier cell lines.
Using an epitope-tagged version of the protein, the truncated protein caused by
c.1592delT was also shown to have markedly decreased BRCA2-binding affinity.
(Erkko et al. 2007, Erkko et al. 2007, Nikkilä et al. 2013.)
2.5.3 Breast cancer predisposition genes of uncertain risk
Additional genes associated with hereditary breast cancer predisposition are
constantly emerging, but how much they contribute to the actual risk of
developing the disease is uncertain. RAD51C is a part of a complex composed of
five RAD51 paralogues (RAD51B, RAD51C, RAD51D, XRCC2, XRCC3)
involved in DNA repair in both the initial and late stages of HR (Somyajit et al.
2010). A recent study identified biallelic mutations in RAD51C that lead to FAlike disorder (Vaz et al. 2010), and this was followed by a study reporting
monoallelic RAD51C mutations associated with increased risk of breast and
ovarian cancer (Meindl et al. 2010). In the initial report by Meindl et al. (2010)
RAD51C germline mutations were identified in 1.3% of families with both breast
and ovarian cancer, but mutation frequency has been varied in subsequent
replication studies performed in different populations. Some follow-up studies
have identified rare deleterious mutations among breast and/or ovarian cancer
families, while others have not found any clearly pathogenic mutations. The level
of penetrance has also varied depending on the RAD51C mutation (Akbari et al.
2010, Levy-Lahad 2010, Meindl et al. 2010, Pelttari et al. 2011, Romero et al.
2011, Thompson et al. 2012, Vuorela et al. 2011, Zheng et al. 2010). Based on the
numerous studies, the current view is that the prevalence of RAD51C mutations,
with the potential exception of founder mutations, is low in the general population.
RAD51C mutations have been speculated to significantly increase the risk of
ovarian cancer, but not breast cancer, in the absence of ovarian cancer family
history, indicating that RAD51C is the first moderate-penetrance susceptibility
40
gene for ovarian cancer. (De Leeneer et al. 2012, Le Calvez-Kelm et al. 2012,
Loveday et al. 2012, Osorio et al. 2012, Pelttari et al. 2012, Walsh et al. 2011.)
The role of the mismatch repair genes (MHL1, MSH2, MSH6, PMS1 and
PMS2) in breast cancer predisposition is still a matter of debate. It is unclear
whether mutations in any of these genes confer to an increased risk to breast
cancer and if that would be the case, how strong the effect would be. They play a
significant role in hereditary non-polyposis colorectal cancer, i.e. Lynchsyndrome (Shanley et al. 2009), but only marginally increase the risk of breast
cancer over the population risk in HNPCC patients (Geary et al. 2008).
RAP80 and ABRAXAS genes occur in a complex with BRCA1 (BRCA1-A)
and they have a critical function in the localizing of BRCA1 to DNA damage foci.
(Akbari et al. 2009, Novak et al. 2009, Osorio et al. 2009). Some studies have
identified rare mutations in both of these genes that predispose to breast cancer
(Nikkilä et al. 2009, Solyom et al. 2012), but these observations have not been
replicated (Novak et al. 2009).
Neurofibromatosis type 1 (NF1) is another debatable gene. NF1 is a common
autosomal dominant disorder characterized by café-au-lait spots, intertriginous
freckling and neurofibromas (OMIM 162200) (Rasmussen & Friedman 2000,
Sorensen et al. 1986). Reports of an association between NF1 and breast cancer
are uncommon. One study suggests that female patients have almost a 5-fold
excessive risk of developing breast cancer by the age of 50 years (Sharif et al.
2007). However, few cases have been reported with a combined diagnosis of NF1
and breast cancer due to mutations in both NF1 and BRCA1 genes in the same
individual (Campos et al. 2013). In conclusion, further population-based studies
are needed to better understand the role of all of these genes in breast cancer
predisposition.
2.5.4 Low-penetrance breast cancer predisposition alleles
Common low-penetrance alleles comprise the third major group of the landscape
of breast cancer susceptibility. They confer very small increases in disease risk,
up to 1.5 RR, but their heterozygote allele frequencies are high, ranging from 10%
to 90% in the general population (Pharoah et al. 2008, Turnbull & Rahman 2008).
The currently known low-penetrance susceptibility alleles were discovered
through association studies, either targeted at individual genes on the basis of
biological candidacy or, more recently (from 2007), through genome-wide tag
SNP searches. A genome-wide association study (GWAS) is a large study that
41
compares the genotype frequencies of hundreds of thousands of common variants
distributed throughout the genome between cases and controls in order to identify
alleles associated with disease risk. GWASs have thus diminished the need for a
good biological understanding of important genes before a disease association can
be detected. This approach has identified common genetic variants that confer
low-penetrance susceptibility to a variety of different cancers, including breast
cancer. (Pennington & Swisher 2012, Trainer et al. 2011)
Common variants at 27 loci have been identified to be associated with breast
cancer susceptibility. In a very recent study, Michailidou et al. (2013) report a
meta-analysis of 9 genome-wide association studies, including altogether 10 052
breast cancer cases and 12 575 controls. SNPs for further genotyping were
collected from 41 different studies in the Breast Cancer Association Consortium
(BCAC). SNPs selected from large multistage GWASs, SNPs selected for fine
mapping of known susceptibility loci, functional candidate SNPs, and SNPs
related to other traits were included in this study. The iCOGS array comprised 211
155 SNPs, and data were obtained for 199 961 SNPs after quality control
exclusion. This study identified SNPs at 41 new breast cancer susceptibility loci,
and based on pre-allele OR estimates these loci explain approximately 5% of the
familial risk of breast cancer. (Michailidou et al. 2013). Most of the breast cancerassociated SNPs identified did not reside in the coding regions, unlike highpenetrance mutations, which would be expected to alter the amino acid sequence
directly or through regulation of splicing. However, some of the SNPs appear to
be associated with an individual gene, many of which have known functions
consistent with a role in breast tumorigenesis, such as growth regulation (FGFR2,
MAP3K1), DNA repair (RAD51L), apoptosis (CASP8), or hormone signalling
(ESR1). For about half of the reported disease risk, associated SNPs occur a long
way from any obvious candidate gene, often in regions described as gene deserts.
However, further investigations of individual loci have provided initial evidence
that these variants may have an effect on long distance chromatin interactions,
which in turn can influence gene-expression several hundred kilobases away from
the variant site. (Easton et al. 2007, Maxwell & Nathanson 2013, McClellan &
King 2010, Pomerantz et al. 2009, Stacey et al. 2007, Trainer et al. 2011.)
The low-risk susceptibility alleles detected through GWASs and targeted SNP
arrays represent an entirely distinct group from high- and moderate-penetrance
alleles. According to the polygenic model, these variants can act synergistically
with environmental or lifestyle factors, and together these variants account for a
modest fraction of familial breast cancer cases. Current studies reveal that the
42
known common low-risk variants constitute ~9% of the familial risk of the
disease. (Latif et al. 2010, Mavaddat et al. 2010, Michailidou et al. 2013,
Turnbull et al. 2010.)
2.6
Susceptibility genes assessed in Finnish clinical genetics
services
Nowadays, when a person with a conspicuous family history of breast and
ovarian cancer comes to a Department of Clinical Genetics for genetic
counselling, the first step is to draw the pedigree and then verify all cancer cases
in the family (to gather the PAD of the tumour and age of diagnosis). If the family
fulfils “the Lund criteria” the probability of finding an inherited mutation in
BRCA genes can be computed by a specific program (Vahteristo et al. 2001). If
this probability is greater than 10%, molecular analysis is offered to the youngest
affected family member as part of genetic counselling. The gene test includes
sequencing and MLPA (multiplex ligation-dependent probe amplification) to
detect large deletions and duplications (Schouten et al. 2002)) for BRCA1 and
BRCA2 genes. If the result is negative for these two susceptibility genes, usually
no further genetic testing for additional rare cancer predisposing syndromes (e.g.
Li-Fraumeni) is offered in Finland, unless indicated by specific findings or family
history (i.e. specific spectrum of different cancer types). Some clinics have
continued gene tests for research purposes and have uncovered mutations in, for
example, CHEK2 and PALB2 genes.
2.7
Searching for novel breast cancer susceptibility genes
Genome-wide linkage analyses using large numbers of families without mutations
in BRCA1 and BRCA2 have not mapped additional susceptibility loci (Smith et al.
2006). This does not completely exclude the existence of further high-penetrance
breast cancer susceptibility genes, but it strongly suggests that, if they exist, they
only account for a very small fraction of familial risk (Stratton & Rahman 2008).
Some years ago, a popular hypothesis was that a large proportion of the familial
cancer risk unaccounted for by known high-penetrace cancer genes would be
explained by common genetic variants of low penetrance (polygenic model).
Interestingly, although GWASs have identified several low-penetrance SNPs, they
explain only a small fraction of familial breast cancer risk (Goldstein 2009). One
explanation for the remaining familial cases is that many uinique or very rare
43
mutations exist with moderate or even high penetrance in a variety of genes
(McClellan & King 2010). In order that the remaining ~70% of familial breast
cancer cases could be explained, further research is necessary.
Disease-causing mutations in moderate-penetrance breast cancer
susceptibility genes do not generally result in large pedigrees with multiple breast
cancer cases, and because their segregation with the disease is incomplete, such
susceptibility genes are not easily mapped by genetic linkage analysis. Moreover,
uncommon disease-causing alleles are unlikely to be detected by association
studies since they exist only in a small number of families. (Stratton & Rahman
2008.) (Figure 4)
Fig. 4. Relation of risk allele frequencies, effect sizes (odds ratios) and feasibility of
identifying risk variants by common genetic techniques. Generally linkage studies are
better for identifying low-frequency alleles with larger effect sizes, whereas
association studies are more effective for identifying common variants with small to
moderate effects. Current evidence suggests that the residual genetic susceptibility is
likely to be due to variants at many loci, each conferring a moderate risk of disease.
Candidate gene approaches are likely to fill a large gap in our knowledge if the genetic
basis of cancer. Modified from (Zemunik & Boraska 2011) and breast cancer
susceptibility genes from (Foulkes 2008). According to present knowledge, common
variants with major effects do not exist in breast cancer.
44
One successful strategy for the identification of additional susceptibility factors is
a candidate-gene approach, i.e. direct interrogation of genes believed to be strong
candidates on the basis of their biological plausibility (Pasche & Yi 2010). Since
most of the moderate-risk genes discovered so far are involved in DNA damage
response, it is important to investigate DNA repair pathways further in order to
find more novel breast cancer susceptibility candidate genes (Meindl 2009,
Ralhan et al. 2007).
Advanced genomic technologies have completely changed the approach to
determining genetic components that constitute traits in breast cancer. Nextgeneration sequencing techniques are continuously being developed. Massive
parallel sequencing can be applied to whole-genome, exome and more targeted
approaches. Exome sequencing focuses on the 1% of the genome that consists of
protein-coding exons and does not require a priori knowledge of which genes are
important. (Pennington & Swisher 2012.)
As certain disease alleles are known to be enriched in isolated populations
(founder mutations), the use of the genetically relatively homogeneous Northern
Finnish population should offer advantages in the search for genes associated
with familial breast cancer. In addition, identification of new inherited breast
cancer susceptibility gene mutations and translation of the research results into
clinical utilization may, in the future, impact patients’ treatment strategy and aid
in the identification of individuals who will selectively benefit from targeted
therapies for their malignancy (Francken et al. 2013).
In 2005, potent PARP [poly (ADP-ribose) polymerase] inhibitors were shown
to selectively inhibit the growth of cells with defects in either BRCA1 or BRCA2
genes (Lord & Ashworth 2012). Later, the same effect was found for PALB2deficient cells (Buisson et al. 2010). The best-understood role of PARP1 protein
is in SSB repair. In normal cells, the effects of PARP inhibition are buffered by
homologous recombination, which repairs the resultant DSB. However, effective
HR is reliant on functioning BRCA1, BRCA2, and PALB2. When these genes are
defective, as they commonly are in tumours of germline mutation carriers, at least
in the case of BRCA1 and BRCA2, DSB are left unrepaired and potent PARP
inhibitors can cause cell death. This gave rise to a new use for these agents; they
are selectively tumour-cytotoxic in BRCA1, BRCA2, and PALB2 mutation carriers.
(Bryant et al. 2005, Farmer et al. 2005, Fong et al. 2009, Lord & Ashworth 2012,
Tutt et al. 2010.)
45
2.8
Molecular pathways to maintain cellular integrity
In normal regenerative tissues, cells grow and divide in response to a myriad of
cues. The outer membrane of most cells is in direct contact with the extracellular
matrix, fluid, and neighbouring cells. Regulatory signals also arise from
intracellular sources. The sequential stages of the cell cycle, particularly S-phase
and mitosis, are exquisitely sensitive to damaged chromosomes and to stalled
DNA replication forks. Mutations in cancer genes affect the responses of cells to
changes in their internal and external environments. Different types of DNA
damage are repaired by distinct multiprotein complexes, because the prime
objective for every life form is to deliver its genetic material intact and unchanged
to the next generation. Each of the ~1013 cells in the human body receives tens of
thousands of DNA lesions every day, and if these are not repaired correctly they
can lead to mutations that can threaten cell viability. (Bunz 2008, Jackson &
Bartek 2009, Lindahl & Barnes 2000.)
2.8.1 DNA damage response pathway
A functional DNA damage response pathway is absolutely essential for cell
viability. Different types of DNA damage are SSBs and DSBs, DNA-protein
cross-links, and damaged bases. Diverse pathways function to repair different
types of damage. In mammalian cells, at least four main pathways of DDR exist:
nucleotide excision repair (NER), base excision repair (BER), homologous
recombination (HR) and non-homologous end-joining (NHEJ). (Bekker-Jensen &
Mailand 2010, Hoeijmakers 2001.) The DDR mechanisms encompass pathways
of DNA repair, which are relatively lesion-specific. NHEJ and HR are the most
important pathways for solving the DSB problem. The repair of DSBs can also
include alternative-NHEJ and single-strand annealing (SSA) (Ciccia & Elledge
2010, Price & D'Andrea 2013). NHEJ, which is less accurate than HR, has a
dominant role in the G1 phase of the cell cycle, when an intact sister chromatin
template is not available and the broken DNA ends are simply ligated together in
an efficient but error-prone fashion. More precise error-free HR is used
predominantly in late S phase and G2, concomitantly with the appearance of the
sister chromatid as a template for the repair. (Downs et al. 2007, Harper &
Elledge 2007, McKinnon & Caldecott 2007.)
Central to the signal transduction pathways are two phosphatidylinositol 3kinase-like kinases (PIKKs), which transmit the damage response signal through
46
phosphorylation. These signal transduction pathways can activate cell cycle
checkpoint arrest or apoptosis. Checkpoints occur at the entry into S phase (the
G1/S checkpoint), at the entry into mitosis (the G2/M checkpoint) and during
replication (intra-S-checkpoints), thus preventing duplication and segregation of
damaged DNA. (Abraham 2004, Kurz & Lees-Miller 2004, Lukas et al. 2004,
Shiloh 2003.) Proteins responsible for DSB repair can be categorized into four
different groups: DNA damage sensors (MRN complex, KU70/KU80, RPA),
transducers (DNA-PK, ATM, ATR), mediators (MDC1, 53BP1, BRCA1,
TOPBP1), and effectors (CHK2, CHK1) (Polo & Jackson 2011).
2.9
Fanconi anaemia
Fanconi anaemia (FA) is a rare recessive genetic disorder associated with a high
frequency of haematological abnormalities, congenital anomalies, and a
predisposition to a variety of cancers (Kee & D'Andrea 2012). FA occurs in about
one in every 100 000 births (Rosenberg et al. 2011). To date, 16 FA genes have
been identified and many more interacting genes discovered (Bogliolo et al.
2013). The majority of FA genes are located on autosomes, with the exception of
FANCB, which is on the X chromosome. Biallelic mutations in any of these FA
genes share a characteristic clinical and cellular phenotype of hypersensitivity to
crosslinking agents and will lead to bone-marrow failure and susceptibility to
both acute myeloid leukaemia (AML) and solid tumours. FA proteins work in
concert in a common cellular pathway, termed the Fanconi anaemia pathway.
These proteins participate in the repair of extraordinarily deleterious lesions and
interstrand crosslinks, but also in the maintenance of genomic stability during
DNA replication. (Kee & D'Andrea 2012, Kottemann & Smogorzewska 2013,
Patel & Joenje 2007, Schlacher et al. 2012.)
The FA core complex is composed of eight upstream FA proteins (FANCA,
FANCB, FANCC, FANCE, FANCF, FANCG, FANCL, and FANCM), which
form a nuclear complex together with accessory proteins (e.g. FAAP20, FAAP24,
and FAAP100). This complex possesses intrinsic E3 ubiquitin ligase activity.
(D'Andrea 2010.) The primary function of the FA core complex is to
monoubiquitinate two FA proteins, FANCD2 and FANCI, after which
monoubiquitylated D2-I heterodimer is recruited to sites of DNA damage by
FANCJ. These FA-proteins localize to nuclear foci and are functionally associated
with the downstream repair factors (Figure 5). The downstream components
consist of five different proteins: BRCA2/FANCD1, RAD51C/FANCO,
47
PALB2/FANCN, BRIP1/FANCJ, and BRCA1. These are dispensable for
monoubiquitination of the I-D2 complex. FAN1 nuclease is recruited and SLX4
(also known as FANCP) functions as a scaffold for the three nucleases XPFERCC1, MUS81-EME1, and SLX1 that function at the site of DNA damage to
make an incision on either side of two covalently linked nucleotides. SNM1A
have a role in processing the crosslink after incision. On the incised strand, TLS
polymerases are recruited to bypass the unhooked crosslink. The break is then
repaired through homologous recombination involving the FA proteins BRCA2,
BRIP1, PALB2, and RAD51C. In addition to these FA or FA-associated proteins,
the FA pathway network broadly includes other regulatory proteins. The
ubiquitin-specific peptidase 1 (USP1) and the USP1-associated protein (UAF1)
regulate deubiquitination of FANCD2/FANCI and are required for completion of
the FA pathway. (Cohn et al. 2007, Kim et al. 2009, Kim et al. 2011, Knipscheer
et al. 2009, Mirchandani & D'Andrea 2006, Niedernhofer et al. 2005, Niedzwiedz
et al. 2004, Prasher et al. 2005, Stoepker et al. 2011.)
Newly diagnosed FA patients should be subtyped because of clinical
variability among subtypes. Mutations in the eight upstream genes are found in
approximately 90% of patients (Kee & D'Andrea 2012). Identification of the
breast cancer susceptibility gene BRCA2 as FA gene FANCD1 (Howlett et al.
2002) reaffirms the close cooperative relationship between the FA pathway and
the breast and ovarian cancer susceptibility BRCA proteins. These tumoursuppressive proteins, BRCA1 (Garcia-Higuera et al. 2001) and BRCA2, act
together with other FA proteins in the repair of DNA interstrand cross-links.
Subsequent identification of FANCN (also known as the partner and localizer of
BRCA2 [PALB2]) (Reid et al. 2007, Xia et al. 2007), FANCJ (also known as
BRCA1-interacting helicase 1 [BRIP1/BACH1] (Levitus et al. 2005, Litman et al.
2005), and more recently, FANCO (RAD51C) (Meindl et al. 2010, Vaz et al. 2010)
further solidify the close association of breast and ovarian susceptibility genes
with FA (Kee & D'Andrea 2012). The current view is that monoallelic mutations
in the four downstream members confer susceptibility to breast and ovarian
cancer, whereas biallelic mutations in the same genes result in FA (Kottemann &
Smogorzewska 2013).
2.9.1 DNA crosslink repair by the Fanconi anaemia pathway
DNA interstrand cross-links (ICLs) are extremely deleterious lesions that
covalently tether both duplex DNA strands and pose formidable blocks to DNA
48
metabolism. Crosslinked DNA impedes both transcription and replication, and
therefore, needs to be removed during all stages of the cell cycle. The FA pathway
is not constitutively active in normal cells, but is turned on during the S phase or
following DNA damage (Mirchandani & D'Andrea 2006, Taniguchi et al. 2002).
The major DNA repair pathway regulated by FA proteins is homologous
recombination (HR), a process that repairs DNA DSBs. The FA proteins promote
HR activity by different mechanisms, and cells lacking FA proteins are deficient
in promoting HR activities. ICL repair requires the tight coordination of several
different repair pathways, and it is within this delicately balanced system that the
FA pathway has its regulatory role. Recent studies have demonstrated how the FA
pathway coordinates three critical DNA repair processes: nucleolytic incision,
translesion DNA synthesis (TLS), and HR (Kim & D'Andrea 2012). Downstream
FA proteins play an important role in HR, but the upstream proteins are also
speculated to be required for HR, although having more minor role (Kottemann &
Smogorzewska 2013, Stecklein & Jensen 2012, Kee & D'Andrea 2012).
The presence of each FA protein is a prerequisite for proper function and
subsequent DNA repair. For example, cells deficient in any of the RAD51
paralogues are sensitive to ICLs and DSBs because of the resulting deficit in
homologous recombination (Huang & Li 2013). However, several FA-interacting
proteins play non-canonical roles also outside of the classical pathway, which
could explain why mutations within different FA genes yield similar, albeit not
necessarily identical clinical phenotypes (Romick-Rosendale et al. 2013).
49
Fig. 5. A schematic view of the Fanconi Anaemia pathway activation for ICL repair.
ATR-mediated signal activates the FA pathway at the stalled replication forks, leading
to the activation of the FA core complex and monoubiquitination of the FANCD2/FANCI
heterodimer. The monoubiquitinated D2/I heterodimer recruits FAN1 nuclease and
FANCP (SLX4), which associates with several nucleases and then coordinates the
action of downstream repair factors. The break is then repaired through homologous
recombination involving the FA proteins BRCA2, BRIP1, PALB2, and RAD51C. Data
combined from (Kee & D'Andrea 2012, Kitao & Takata 2011, Soulier 2011).
2.9.2 FA genes; FANCA, FANCG and FANCI
FANCA and FANCG are part of the FA core complex. As described earlier
(Section 2.8.2), the primary function of the FA core complex is to
monoubiquitinate two FA proteins, FANCD2 and FANCI, which are expressed as
a constitutive heterodimer (Crossan & Patel 2012). (Table 1)
50
Table 1. Three FA genes.
Gene
FANCA
FANCG (XRCC9)
FANCI
Locus
Mutation
Amino
Protein
frequency
acids
MW (kDa)
Notable protein function
16q24.3
~66%
1455
163
9p13
~10%
622
70
Scaffold for FA core complex
15q25-26
<2%
1328
150
Monoubiquitinated, forms
Scaffold for FA core complex
heterodimer with FANCD2
FA-A is the major FA complementation group and the corresponding gene FANCA
was cloned in 1996 (Fanconi anaemia/Breast cancer consortium 1996, Lo Ten Foe
et al. 1996). It is a large gene consisting of 43 exons. Mutation analysis in FA-A
patients has not revealed any mutational hotspots (Savino et al. 1997, Wijker et al.
1999). Fanca-/- cells presented an increase in spontaneous chromosomal damage
(22% vs. 4% in wild-type) when MMC was absent (Cheng et al. 2000).
FANCG localization was achieved by homozygosity mapping and linkage
analysis in FA-G families. FANCG covers 14 exons and the protein shows no
homology to known proteins. (Kruyt et al. 1999, Saar et al. 1998, Waisfisz et al.
1999.) Fancg knock-out cells have not only decreased HR capacity for repairing
enzymatically induced, site-specific chromosomal DSBs but also increased
ionizing radiation (IR)-induced chromosomal aberrations in cells irradiated in late
S and G2 phases (Yamamoto et al. 2003). The overall structure of the FANCG
protein seems to be critical for the interaction. FANCG has been shown to interact
with the RAD51 paralogue protein XRCC3 in the yeast two-hybrid system.
(Hussain et al. 2006.)
FANCI has 38 exons and its protein contains three nuclear localization and
three predicted ATM/ATR phosphorylation motifs. Patient cell lines of
complementation group I are deficient in FANCD2 monoubiquitination, show
increased chromosomal instability and are hypersensitive to cross-linking agents
such as mitomycin C (Dorsman et al. 2007). The phosphorylation of FANCI has
been suggested to function as a molecular switch to turn on the FA pathway, but
further investigations are required to confirm this (Ishiai et al. 2008). In addition,
in FANCA- and FANCG-deficient cell lines, FANCI monoubiquitination was
completely abrogated, in contrast, FANCI ubiquitination was restored in
complemented cells, that contained an intact FA core complex. (Leung et al. 2012,
Mirchandani & D'Andrea 2006.)
51
2.10 Candidate genes for breast cancer predisposition based on
their biological function
Given the numerous breast cancer predisposition genes that are part of the
BRCA/FA signalling machinery and involved in genomic integrity maintenance,
all related factors are good candidates for breast cancer susceptibility genes, until
disproven. In addition, besides the FA genes, all of the other genes involved in
other critical biochemical pathways, such as signal transduction pathways that
transmit checkpoint signals in response to DNA damage, replication blocks and
spindle damage, have been proposed as candidates for cancer susceptibility.
(Elledge 1996, Nathanson & Weber 2001, Ralhan et al. 2007, Yoshida & Miki
2004.) The candidate gene approach involves the systematic screening of entire
genes for disease-causing variants in cancer cases, and the frequency of the
observed mutations is compared with that of controls. The genes are chosen on
the basis of a priori knowledge of their biological activity. The usage of familial
rather than population-based breast cancer cases is expected to increase the power
for mutation detection (Antoniou & Easton 2003).
2.10.1 AATF
AATF (apoptosis-antagonizing transcription factor, also known as CHE1 and
Traube) is an evolutionarily conserved RNA polymerase II binding protein
involved in the regulation of gene transcription. The AATF gene is located at
chromosome 17q11.2-q12. It encodes a phosphoprotein containing 558 amino
acids and consists of 12 exons. (Fanciulli et al. 2000)(Lindfors et al. 2000). It
contains a leucine zipper motif, several phosphorylation sites for different kinases,
a nuclear localization signal motif, three nuclear receptor binding motifs, and
several four-nucleotide repeats (Lindfors et al. 2000, Passananti et al. 2007).
The functions of AATF are essential for early embryogenesis and cell
proliferation (Bruno et al. 2002, Thomas et al. 2000). One significant function of
AATF is to promote cellular transcription; it acts as an adaptor that links specific
transcription factors to the general transcription apparatus. Furthermore, due to its
interaction with various important components of the cell survival machinery, e.g.
anti-apoptotic factors, guch as XIAP, and pro-apoptotic factors, such as Dlk, Par-4,
and NRAGE, AATF has been found to play an important role in DNA damage
response, cell cycle checkpoint control, and also chromatin remodelling (Jagut et
al. 2013, Kaul & Mehrotra 2007, Passananti et al. 2007). In addition to pro52
proliferative function, AATF exhibits strong anti-apoptotic activity, and this
protein is downregulated during apoptosis (De Nicola et al. 2007, Guo & Xie
2004, Page et al. 1999). In response to DNA damage, AATF relocates to the TP53
promoter to increase transcription of this gene, thereby contributing to the
increase in p53 protein levels after DNA damage (Bruno et al. 2006). The p53
protein plays a critical role in the cellular response to DNA damage and other
stresses by inhibiting proliferation or by inducing apoptosis. The p53 is the most
frequent target for genetic alterations in human cancer. (Bruno et al. 2010,
Hopker et al. 2012.)
Centrosomes are part of a network that integrate cell cycle arrest and repair
signals. Crucial regulators of DDR, such as the checkpoint kinases ATM, ATR,
Chk1, and Chk2 as well as the tumour suppressors p53 and BRCA1, have been
found to localize to the centrosomes (Ciciarello et al. 2001, Oricchio et al. 2006).
Centrosomes integrate G2/M checkpoint control and repair signals in response to
genotoxic stress (Zhang et al. 2007). Depletion of AATF prevents centrosomal
recruitment of the checkpoint kinase Chk1 and leads to abnormal centrosome
amplification, accumulation of multinucleated cells and abnormal spindle
formation. In conclusion, AATF is critical at the spindle assembly checkpoint.
(Sorino et al. 2013.)
2.10.2 MRG15
Human MRG15 (MORF4-related gene on chromosome 15) was first identified as
a member of the MORF4/MRG family of novel transcription factors. Of seven
known family members, only three are expressed (MORF4, MRG15, and MRGX).
MORF4 induces senescence in a subset of human tumour cell lines, and it is a
truncated version of MRG15. In contrast to MORF4, MRG15 and MRGX are
positive regulators of cell division. MRG15 is the largest of these three, encoding
a 323 amino acid protein. MRG15 has an ATP-GTP binding region, followed by a
helix-loop-helix and a leucine zipper at the C terminus. MGR15 also encodes a
bipartite nuclear localization signal (NLS) flanked by phosphorylation sites.
MRG15 is broadly evolutionarily conserved from yeast to mammals and even
plants, suggesting it must have a fundamental activity in some cellular process.
(Bertram & Pereira-Smith 2001, Chen et al. 2010, Pena & Pereira-Smith 2007,
Tominaga & Pereira-Smith 2002, Tominaga et al. 2005.)
MRG15 protein specifically binds to Lys36-methylated histone H3 (Zhang et
al. 2006) and has an established role in transcriptional regulation (Carrozza et al.
53
2005). In particular, MRG15 and MRGX are integral components of the
NuA4/Tip60 histone acetyltransferase (HAT) and histone deacetylase (HDAC)
complexes (Doyon et al. 2004, Hayakawa et al. 2007). Several lines of evidence
suggest that MRG15 and its related proteins are involved in DNA-repair
processes (Garcia et al. 2007, Nakayama et al. 2003). MRG15 binds directly to a
key component of the machinery for DSB repair, PALB2 (Sy et al. 2009a).
Furthermore, MRG15 is required for HR and resistance to MMC. MRG15
participates in the response to DSBs by recruiting the BRCA complex to sites of
damaged DNA (Hayakawa et al. 2010). Results in human, mouse, and C. elegans
models delineate molecular and functional relationships with BRCA2, PALB2,
RAD51, and RPA1 (Takashi et al. 2009), suggesting a role for MRG15 in the
repair of DNA DSBs. An MRG15 mutant lacking the C-terminal chromo domain
was unable to interact with PALB2. Furthermore, a Mrg15-deficient murine
embryonic fibroblast showed reduced formation of Rad51 nuclear foci. (Martrat
et al. 2011.)
Based on the role that both these genes AATF and MRG15 have in DDR, it is
possible that their mutations could predispose to cancer.
54
3
Aims of the study
The currently identified breast cancer susceptibility genes only account for about
25% of familial breast cancer patients. Thus, the vast majority of cases cannot be
explained by any known predisposing factors. Numerous additional predisposing
genes are speculated to await discovery. The majority of known susceptibility
genes have been reported to be involved in the maintenance of genomic stability
via the DNA damage response. Other genes involved in the same pathways may
therefore be good candidate genes for familial breast cancer susceptibility. In this
work, the AATF and MRG15 genes were evaluated as potential candidates as both
are known to act in critical cellular pathways. Moreover, PALB2 and other
Fanconi anaemia (FA) genes are previously known to be breast cancer
susceptibility genes but their clinical significance has not been thoroughly
investigated in the Finnish population.
Specific aims of the studies are as follows:
1.
2.
3.
4.
To identify potential cancer-predisposing mutations in the AATF gene and to
evaluate their involvement in breast cancer susceptibility.
To screen MRG15 for conventional mutations and to assess their potential
role in breast cancer predisposition.
To evaluate whether Finnish PALB2 c.1592delT founder mutation testing
should be part of routine clinical counselling by investigating a cohort of
high-risk breast cancer patients at the Department of Clinical Genetics, Oulu
University Hospital. The patients had previously tested negative for BRCA1
and BRCA2 mutations. Specifically, the aim was to determine whether PALB2
c.1592delT mutation analysis would give the patients some direct additional
advantage if this test was included in routine aetiological search of clinical
determinants for familial breast cancer susceptibility.
To identify all current Finnish FA patients and to determine their diseaseassociated complementation groups and the causative mutations, and further,
to investigate the putative role of these mutations in breast and prostate
cancer susceptibility.
55
56
4
Materials and methods
4.1
Subjects
4.1.1 Study I
Index patients from 121 breast or breast and ovarian cancer families from
Northern Finland were included in the mutation screening of the AATF gene. For
statistical purposes, one index patient from each family was chosen according to
the youngest age of breast cancer onset. Mutations were screened in the coding
regions and the exon-intron flanking sequences of the AATF gene. The families
derived geographically from the Northern Ostrobothnia Health Care District, and
were categorized as high- and moderate-risk families. The inclusion criteria for
high-risk families were as follows:
1.
2.
Three or more breast and/or ovarian cancer cases in first and second-degree
relatives.
Two cases of breast and/or ovarian cancer in first- and second-degree
relatives, of which at least one with early age of disease onset (<35 years),
bilateral disease or multiple primary tumours.
The remaining families were considered moderate-risk families with two cases of
breast or breast and ovarian cancer in first- or second-degree relatives. The total
number of high-risk families was 71 and moderate-risk families 51.
All of the high-risk families were previously screened for germline mutations
in BRCA1, BRCA2, CHK2, and TP53 (Allinen et al. 2001, Huusko et al. 1998,
Huusko et al. 1999, Rapakko et al. 2001). In addition, the families had been
screened for mutations in RAD50, MRE11, ATM, NBS1, BRIP1/BACH1 and
BARD1 genes (Heikkinen et al. 2003, Heikkinen et al. 2005, Karppinen et al.
2003, Karppinen et al. 2004).
The healthy control samples were from consecutive anonymous (only gender,
age, and place of blood donation were known) Finnish Red-Cross blood donors
originating from the same geographical region as the familial breast cancer cases.
All control individuals were cancer-free at the time of blood sample donation.
Donor health status was not followed. The age of the blood donors varied
between 18 and 65 years, the average being 42 years.
57
4.1.2 Study II
The patient material in this study was the same as in Study I, with the execption
that the number of analysed Northern Finnish breast cancer patients was higher,
148. This included 99 high-risk and 49 moderate-risk families. Study II also
included 84 index patients of breast/ovarian cancer families originating from
Canada. Of these, 45 were French-Canadian and 9 Ashkenazi Jewish families.
The total number of families analysed was 232. The average age at diagnosis for
the index cases was 48 years, and 97% of probands presented with breast cancer.
The classification and the actual number of participating patients were as follows:
1.
2.
3.
179 (77%) of the 232 families had three or more affected individuals.
44 families (19%) had two affected individuals.
Nine index cases (4%) presented with either breast cancer at age under 30
years or bilateral cancer, but with no family history.
4.1.3 Study III
Subjects were selected among the individuals (n=223) who had been in contact
with the Oulu University Hospital, Department of Clinical Genetics, between
1997 and 2011. Inclusion to the study was in accordance with the following
criteria of high-risk hereditary breast cancer (“the Lund criteria”):
1.
2.
3.
4.
5.
The individual or her first-degree relative (only female family members were
included when defining first-degree relatives) had breast and/or ovarian
cancer at age under 30 years.
Two first-degree relatives in the family had breast and/or ovarian cancer and
at least one of the cancers had been diagnosed at age under 40 years.
Three first-degree relatives in the family had breast and/or ovarian cancer and
at least one of the cancers had been diagnosed at age under 50 years.
Four or more relatives had breast and/or ovarian cancer.
The same individual had both breast and ovarian cancer. Individuals with
bilateral breast cancer were considered to have two separate cancers.
BRCA1- or BRCA2-positive mutation status was an exclusion criterion; 16
individuals were excluded based on this. Of the initial 223 individuals, 101 met
the inclusion criteria based on their pedigree data; 10 were deceased and the
remaining 91 were invited to participate. Seventy individuals responded (70/91 =
77%) and provided us with information on their family history of cancer. Chart
58
review led to the further exclusion of 14 individuals who did not meet the
selection criteria. Altogether, the Northern Finnish 1997–2011 study cohort tested
for PALB2 c.1592 carrier status consisted of 62 (56 alive and 6 deceased, who had
participated in one of our previous studies on PALB2 (Erkko et al. 2007)) affected
females (Figure 5).
10 were deceased
but 6 of them had
been tested for PALB2 mutation previously 233 individuals
contacted the genetic
counselling unit
because of suspected
breast/ovarian cancer
101 had a major
family history of cancer and were
negative for mutations
in BRCA1/BRCA2
122 did not meet study criteria
70 of 91 individuals
responded
to the study invitation
Altogether 62 eligible
index patients were
analysed for
PALB2
mutation status
Chart review of updated family
history data excluded a further 14 individuals
Fig. 6. Flow chart summarizing the numbers of patients at different stages of the
inclusion process in Study III.
4.1.4 Study IV
An attempt to collect all known Finnish FA patients was performed by contacting
clinicians (clinical genetic/paediatrician) to inquire about patients with suspected
or confirmed diagnosis of FA. In addition, all genetic laboratories were contacted
to find out if they had performed chromosomal analysis for suspected cases. The
patients were collected as a part of a collaborative study. Three patients had been
referred for FA clinical testing e.g. diepoxybutane (DEB) and mitomycin C
(MMC) sensitivity test by Pediatric Oncology, Tampere University Hospital and
one by Family Federation of Finland. Additionally, five anonymous cell lines
created during the years 1979–1993 from individuals with suspected FA
(Department of Clinical Genetics, Helsinki University Central Hospital) were
collected. These cell lines were subjected for MMC sensitivity testing by cell
survival assay and only one showed positive and was included in the study.
59
The Northern Finnish familial cohort consisted of 164 index cases from
breast or breast-ovarian cancer families. The inclusion criteria were the same as in
Study I. An additional young breast cancer cohort was included, consisting of 82
cases from Northern Finland who were diagnosed with breast cancer at an age of
≤40 (from 25 to 40 years, with median 38 years). These young patients were
unselected for a family history of the disease. The prevalence of FA mutations
was tested also in the Northern Finnish breast cancer cases who had been
operated on at Oulu University Hospital in 2000–2007 (n=508–813, depending on
the tested mutation). These patients were unselected for age at diagnosis as well
for family history of breast cancer. The Northern Finnish control samples (n=517)
were derived from anonymous cancer-free female Finnish Red Cross blood
donors who had originated from the same geographical region as the patient
cohorts.
The Southern Finnish familial cohort included index cases from 87 hereditary
breast and ovarian cancer families from the Pirkanmaa Health Care District. All
of these individuals had been founder mutation-negative by minisequencing of 28
known Finnish BRCA1/BRCA2 mutations and by protein truncation test of exon
11 for BRCA1 and exons 10 and 11 for BRCA2. The samples had been collected
from individuals who had visited the Tampere University Hospital Genetics
Outpatient Clinic between January 1997 and May 2008. The Southern Finnish
unselected breast cancer cohort included 694 breast cancer cases unselected for
family history of cancer and was recruited through the Tampere University
Hospital. Samples from 659 anonymous ancestry matched healthy female Finnish
Red Cross blood donors were used as controls.
The Eastern Finnish cohort (Kuopio University Hospital), comprising 494
patients with breast cancer from the Kuopio Breast Cancer Project (KBCP), was
utilized for genotyping of FANCG mutations. These patients originated from the
Northern Savo region and were diagnosed at Kuopio University Hospital in
1990–1995. Serving as controls were 468 female subjects selected from the
National Population Register. This additional cohort was available only for
FANCG mutations genotyping.
The prostate cancer cohort included 565 prostate cancer index patients from
Southern Finland (Pirkanmaa Health Care District). All familial prostate cancer
cases came from families with two or more affected members (the youngest
affected male from each family was included in the analysis). Serving as controls
were 469 unaffected males from the same geographical region.
60
4.2
DNA extraction
Genomic DNA from peripheral blood lymphocytes was extracted by a standard
phenol-chloroform method or by using the Purgene D-50K purification kit
(Gentra, Minneapolis, MN, USA).
4.3
Mutation detection methods
4.3.1 Conformation-sensitive gel electrophoresis (I)
Mutation screening of the AATF was done using conformation-sensitive gel
electrophoresis (CSGE). This method was used to scan the coding regions and
exon-intron boundaries of candidate genes. The assay is based on denaturating gel
separation of amplicons with conformation changes due to mismatches in the
double-stranded DNA. This results in different migration patterns between homoand heteroduplexes (Ganguly et al. 1993, Körkkö et al. 1998).
4.3.2 High-resolution melt analysis (II, III, and IV)
Mutations in the protein encoding regions and exon-intron boundaries of the
candidate genes were searched for by high-resolution melting (HRM) analysis.
This is a post-PCR analysis method for identifing nucleic acid sequences that is
based on detecting small differences in PCR melting (i.e. dissociation) curves.
Wild-type and heterozygous samples can be differentiated in the melting plots.
(Reed & Wittwer 2004.)
4.3.3 Direct sequencing (I, II, III, IV)
Samples with band shifts in CSGE or deviant melting curves in HRM (I, II, III)
were reamplified and sequenced with the Li-Cor IR2 4200-S DNA Analysis
system (Li-Cor Inc., Lincoln, NE, USA) or with capillary sequencing using
ABI3130xl Genetic Analyzer (Applied Biosystems, Foster City, CA, USA).
Sequencing was also used as a primary method to detect mutations in sequences
not suitable for analysis with CSGE or HRM. In study IV, after identification of
the complementation groups, the underlying gene mutations of each patient were
studied further by sequencing. For Li-Cor, the Sequi Therm EXEL II DNA
61
Sequencing kit-LC (Epicentre Technologies, Madison, WI, USA) and for ABI the
Big Dye Terminator kit v1.1 (Applied Biosystems) were used.
4.4
Mitomycin C sensitivity test and complementation group
analysis (IV)
Based on the hypersensitivity of FA cells to ICL, the FA status of clinically
suspected patients was confirmed with chromosomal dieposybutane (DEB)
mitomycin C (MMC) sensitivity test from peripheral blood lymphocytes (German
et al. 1987).
All the confirmed FA cases were subjected to complementation group
analysis (Chandra et al. 2005). For FA complementation group determination,
five patient cell lines were established. Cells were transduced with retroviral
vectors containing normal cDNAs of FANCA, FANCB, FANCC, FANCE,
FANCF, FANCG, or FANCL and analysed for cell cycle arrest upon treatment
with MMC and for restoration of FANCD2 monoubiquitination. (Hanenberg et al.
2002.)
4.5
Statistical and bioinformatical analysis
Carrier frequencies between patients and healthy controls were compared using
Pearson’s Chi-Square or Fisher’s exact test (I, II, IV) with SPSS version 16.0–
20.0 for Windows (SPSS Inc., Chicago, USA). All p-values were two-sided.
In Studies I and II, all alterations were tested for possible pathogenicity by
PolyPhen software (Ramensky et al. 2002) and were checked with NNSplice
software for potential effects on splicing (Reese et al. 1997) (Reed & Wittwer
2004)ESEfinder 2.0 software (Cartegni et al. 2003) was used to determine
whether the exonic variant was located in ESE (exonic splicing enhancer)
sequences and could therefore influence in ESE function.
For Study II, also in silico analyses were performed using accurate splicing
prediction software: BDGP (Reese et al. 1997), NetGene 2 server (Brunak et al.
1991, Hebsgaard et al. 1996), and GENESCAN (Burge & Karlin 1997). Analysis
of the 3’UTR was performed by using TargetScanHuman 5.1 software (Lewis et
al. 2005).
62
4.6
Ethical issues
Investigations aiming to find new breast cancer susceptibility genes may lead to
discoveries which improve genetic counselling, help early diagnostics in relatives,
improves understanding of breast cancer and may eventually lead to new
treatment for breast cancer. Thus this study has an ethically sound aim.
All participating individuals gave informed consent for using their pedigree
data and blood specimens in a study on cancer susceptibility gene mutations.
Each DNA sample was labelled with a research code, and samples were handled
anonymously during the analyses. Results from the analyses in Studies I, II, and
IV were only used for research purposes, and no information was given to the
participating individuals. In contrast, in Study III a separate invitation to
participate in the study was issued and a new informed consent obtained.
Participants had the opportunity to choose whether they wanted to know their
mutation status or not, and those requesting the information were informed by a
personal letter. Genetic counselling was provided, if desired, for identified
mutation carriers.
Appropriate research permissions to perform the study have been obtained
from the Ethics Committee of Oulu University Hospital (88/2000 and 12/2011) as
well as from other participating University Hospitals. The study protocol was
approved by the Ministry of Social Affairs and Health (46/07/98).
63
64
5
Results
5.1
Mutation screening of the AATF gene (I)
Screening for germline mutations in AATF in 121 index patients with
breast/ovarian cancer revealed altogether seven different changes (Table 2). Only
one of the observed changes was exonic. The alteration c.739G>T in exon 4
resulted in an Ala247Ser amino acid substitution. All other variants were intronic.
Of the total of seven changes, only two were previously reported; five were novel.
To evaluate possible pathogenicity of the observed changes, their frequencies
were compared between cases and healthy population controls. The novel exonic
alteration was observed at similar frequencies in both cases and controls (p=0.70).
Although it resulted in an amino acid change, PolyPhen software analysis
predicted the effect to be neutral. Details and occurrence of the observed
nucleotide variants are shown in Table 2.
Table 2. Observed sequence variations in the AATF gene.
Controls
P-value (OR, 95% CI)
Previously
Exon/
Nucleotide
Effect on
Carrier
Intron
change
protein
frequency (%)
known (+) or
Familial cases
unknown (-)
alteration
Ex4
c.739G>T
p.Ala247Ser
1.7 (2/121)
1.3 (4/317)
Int3
c.694+35A>C
None
4.1 (5/121)
9.1 (30/328) 0.08 (0.4, 0.2–1.1)
c.694+48T>G
None
3.3 (4/121)
0.9 (3/328)
0.09 (3.7, 0.8–16.8)
-
c.832+17C>T
None
0.8 (1/121)
0.3 (1/317)
0.48 (2.6, 0.16–42.4)
-
c.832+39C>T
None
0.8 (1/121)
- (0/317)
Int11 c.1619+29A>C
None
2.5 (3/121)
0.3 (1/325)
Int12 c.1670+42C>T
None
20.7 (25/121)
Int4
5.2
0.70 (1.3, 0.2–7.1)
+; rs8067751
0.28 (NA)
-
0.06 (8.2, 0.8–80.0)
-
23.5 (72/307) 0.53 (0.9, 0.5–1.4)
+; rs11653434
Mutation screening of the MRG15 gene (II)
In the MRG15 gene, altogether 21 variants were identified in affected index
individuals from 232 breast or breast/ovarian cancer families (Table 3). Of these
variants, 17 were intronic changes, 3 were coding SNPs, and 1 was localized in
the 3’UTR. Six of the 17 intronic variants as well as the variant located in the
3’UTR were novel, and they were not referenced in either dbSNP or Ensembl
SNP databases. Each of these seven variants was detected in less than 1% of the
65
patients, suggesting that they were rare and potentially harmful cancer-associated
variants.
To gain further insight into possible disadvantageous effects of these seven
non-referenced variants, in silico analysis using accurate splicing prediction
BDGP, NetGen 2 server, and GENESCAN software was performed. None of
these intronic changes were, however, predicted to impact the splicing of premRNA. Further, analysis of the 3’UTR by using TargetScanHuman 5.1 software
revealed that the site where this variant resides is not a known target of the
miRNAs. Consequently, it is very unlikely that this variant has any effect on the
regulation of MRG15 mRNA expression.
Table 3. Observed sequence changes in the MRG15 gene.
Sequence variant
Protein change
No. of times observed Additional information
c.421+35C>T
Int4-5
7
rs16970187
c.590-26T>C
Int6-7
3
rs4778952
c.616-79T>A
Int7-8
18
rs10519215
c.721A>G
N191S
1
rs73477676
c.807-144T>C
Int9-10
1
Not referenced
c.807-133T>C
Int9-10
1
Not referenced
c.824C>A
A225A
18
rs1435163
c.868A>G
K240R
18
rs61734867
c.895+32A>G
Int10-11
1
rs74024473
c.895+38A>G
Int10-11
1
rs56816206
c.1068+12A>G
Int11-12
2
rs16970207
c.1069-45C>A
Int11-12
1
Not referenced
c.1153+95T>A
Int12-13
11
rs10519216
c.1154-185T>A
Int12-13
1
Not referenced
c.1154-43T>C
Int12-13
19
rs12438433
3’UTR
1
Not referenced
c.1238+161G>C
5.3
Clinical significance of PALB2 c.1592delT screening in BRCAnegative high-risk breast cancer families (III)
Altogether 56 living breast cancer patients who tested negative for BRCA1 and
BRCA2 mutations were included in the PALB2 c.1592delT screening study.
Among these index patients, two were found to carry the heterozygous PALB2
c.1592delT mutation (2/56=3.6%). In addition, the mutation status was known for
66
six other previously tested but recently deceased patients. Among these
individuals, one more mutation carrier was observed. Therefore, altogether 4.8%
(3/62) of the individuals belonging to the Northern Finnish 1997–2011 study
cohort and counselled at the Department of Clinical Genetics at Oulu University
Hospital were carriers of the PALB2 c.1592delT allele.
The average age of disease onset among mutation carriers in the current study
was 48 years. Tumour characteristics of PALB2 c.1592delT mutation carriers were
obtained from pathology records (Table 4). All tumours were histological ductal,
and both oestrogen and progesterone hormone receptor statuses were positive.
Table 4. Tumour characteristics and family history of cancer among identified PALB2
c.1592delT mutation carriers.
Family ID Cancer
Histo-
TNM
Er/Pr
Grade
Her2
p53
Ki67 Other cancers in the
logy
family
Fam A
Br(63)
Ductal
T2N0M0 Er+/Pr+
2
neg
Neg
2
Fam B
Br(43)
Ductal
TisN0M0 Er+/Pr+
1
neg
NA
0
Br(44), Br(73)
Br(30), Br(53), Br(65), Br,
Br, Lung, Ov(60), Panc,
Sarc, Skin, Sto
Fam C
Br(39)
Ductal
T2N0M0 Er+/Pr+
3
NA
Neg
2
Br(67), Br(68), Ov(71),
Panc, Ut(36), Ut
Er, oestrogen receptor; Pr progesterone receptor; Her2, human epidermal growth factor receptor 2; p53,
tumour protein 53; Ki67, antigen Ki67; Br, breast cancer; Ov, ovarian cancer; Panc, pancreatic cancer;
Sarc, Sarcoma; Sto, stomach cancer; Ut, Uterine cancer, NA, not available.
5.4
Complementation groups of Finnish Fanconi anaemia patients
and role of causative mutations in breast cancer susceptibility
(IV)
In Study IV, all of the five analysed Finnish FA patients were confirmed to belong
to previously described FA complementation groups: FA-A (three patients), FA-I
(one patient), and FA-G (one patient). In total, six different FA mutations were
identified in these patients (see Table 5)
67
Table 5. Observed mutations among Finnish FA patients.
Complementation Gene mutations
Additional information
group
FA-G
FANCG c.832insG and c.1076+1G>A
compound heterozygous
FA-I
FANCI c.3041G>A* and c.2957_2969del
*rs140404896
compound heterozygous
FA-A
FANCA c.3348-1G>A homozygous
Described in FA patient in compound
heterozygous state (Chandrasekharappa et
al. 2013)
FA-A
FANCA c.3348-1G>A homozygous
FA-A
FANCA c.-4403_1084-126del
Heterozygous mutation reported in familial
homozygous
breast cancer (Solyom et al. 2011)
The mutation in the FA-G patient was compound heterozygous. One allele
showed a c.832insG mutation in exon 7, which leads to a frameshift
(p.Glu275Glyfs) and further to a truncated protein product. The other FANCG
allele harboured a c.1076+1G>A mutation, which affects a splice site of exon 8.
This leads to an in-frame deletion of 18 nucleotides at the mRNA level
(c.1059_1076del) and abolishes 6 amino acids of FANCG’s third TRP domain.
TRPs are 34-amino acid repeat motifs functioning as scaffolds that mediate
protein-protein interactions (Blom et al. 2004).
The FANCI patient was also compound heterozygous with regard to diseaseassociated mutations. FANCI c.2957_2969del (p.Val986Alafs) leads to a truncated
protein product missing part of exon 26 and all of exons 27–38. FANCI
c.3041G>A was previously referenced (rs140404896), and it leads to a
p.Cys1014Tyr missense mutation in exon 27, changing an evolutionarily highly
conserved amino acid. Based on the analysis with PolyPhen, this substitutional
change probably has damaging consequences on protein function with a score of
0.995. SIFT software predicts that it is damaging with a score of 0.02.
All three FA-A patients were homozygous for their mutations. One of them
had a large deletion of FANCA c.-4403_1084-126del (FANCA ex1-ex12del),
which removes the promoter and the first 12 exons of the gene. This allele has
previously been described in a familial breast cancer case and is predicted to
result in a null allele (Solyom et al. 2011). The two other FA-A patients were
siblings and they both had the same mutation. They were homozygous for the
FANCA c.3348-1G>A mutation, which leads to skipping of exon 34 (c.3348_3408)
at the mRNA level and to an in-frame deletion of 20 amino acids.
68
5.4.1 Finnish FA mutations prevalence in cancer cohorts
The association of the observed Finnish FA mutations with breast cancer
susceptibility was studied using familial and unselected breast cancer cohorts.
Altogether two FA mutation carriers were observed in familial cases: one FANCA
ex1-ex12del (1/321, 0.3%) and one FANCI c.2957_2969del (1/333, 0.3%) carrier.
The first one has previously been described in a family where the index case had
breast cancer at the age of 50 years and her father was diagnosed with pancreatic
cancer at the age of 65 years, and they were both carriers of this FANCA ex1ex12del mutation (Solyom et al. 2011). From the unselected patient material,
three more FANCA ex1-ex12del carriers were identified (3/1495, 0.2%). One of
them had breast cancer at the age of 33 years and had a family history of lung
cancer (three first-degree relatives, smoking status unknown). Another unselected
patient had breast cancer at the age of 52 years, and her aunt and uncle were
diagnosed with pancreatic and lung cancer, respectively, but their carrier status
was unknown. The third FANCA ex1-ex12del carrier from the unselected cohort
had breast cancer at the age of 71 years, and no other cancer history was reported
in her family. None of these unselected cases had additional samples from
relatives available for mutation testing. The prevalence of this FANCA ex1ex12del was also tested among healthy controls, and the result was somewhat
surprising. Two mutation carriers were identified (2/1175, 0.17%), but at least one
of them was very young (<25 years) at the time of blood donation.
FANCI mutation c.2957_2969del was observed in a patient with both ovarian
(at 69 years) and breast cancer (at 72 years). The patient’s mother was suspected
of having a benign ovarian tumour, and the mother’s sister and her daughter both
had cervical cancer. Other samples from family members were not available to
test for mutation segregation. Two FANCI c.2957_2969del carriers were also
observed in the unselected breast cancer cohort (2/1191), but also among healthy
controls (2/896).
The remaining FA mutations, FANCA c.3384-1G>A, FANCI c.3041G>A,
FANCG c.1076+1G>A, and FANCG c.832insG, were observed in either
unselected cases or controls, or both, but their individual frequencies remained
very low, and no significant difference emerged between cases and controls. No
FANCG c.1076+1G>A mutations were observed in familial or unselected cases,
but one carrier was identified among the controls (0.09%). In contrast FANCG
c.832insG was observed in one unselected case (0.06%), but not in healthy
controls (table 6).
69
Table 6. FA mutation frequencies in breast cancer cohorts and in healthy female
controls.
FA mutation
Carrier frequencies and statistics
Familial/young
Unselected
Unaffected
breast cancer
breast cancer
females
1/321 (0.3%)
3/1495 (0.2%)
2/1175 (0.2%)
P-value (OR; 95% CI)
F (Familial/Young),
U (Unselected)
FANCA
ex1-12dela
FANCA
0/321 (0%)
1/1495 (0.1%)
1/1175 (0.1%)
c.3384-1G>A
FANCI c.2957_2969del
(F) 0.5 (1.8; 0.1–20.3)
(U) 1.0 (1.2; 0.2–7.1)
(F) 1.0 (NA)
(U) 1.0 (0.8; 0.05–12.6)
1/333(0.3%)
2/1191 (0.2%)
2/896 (0.2%)
(F) 1.0 (1.3; 0.1–14.9)
(U) 1.0 (0.8; 0.1–5.3)
FANCI
0/333 (0%)
1/1191 (0.1%)
2/896 (0.2%)
c.3041G>A
FANCG c.832insG
(F) 1.0 (NA)
(U) 0.6 (0.4; 0.03–4.1)
0/317 (0%)
1/1716 (0.06%)
0/1264 (0%)
(F) 1.0 (NA)
(U) 1.0 (NA)
FANCG c.1076+1G>A
0/317(0%)
0/1714 (0%)
1/1165 (0.09%) (F) 1.0 (NA)
(U) 0.4 (NA)
a
FANCA ex1-12del screening partially reported in Solyom et al. (2011), consisted of 100 familial breast
cancer cases, 540 unselected cases, and 124 controls from Northern Finland. The one mutation-positive
family was identified in the previous study. NA, not available
70
6
Discussion
6.1
No mutations found in the AATF gene among Finnish breast
cancer families (I)
A significant fraction of familial breast and/or ovarian cancer cannot be explained
by mutations in the currently known susceptibility genes. AATF encodes a
phosphoprotein required for maintaining genomic integrity and cell cycle
checkpoint control (Kaul & Mehrotra 2007, Lindfors et al. 2000). Based on its
important biological role, it was hypothesized that mutations in this gene might
predispose to cancer.
To our knowledge, this study is the first to explore the role of AATF in breast
cancer susceptibility. The whole coding region and exon-intron boundaries of the
AATF gene were screened for germline mutations. Although several sequence
variants were found, none are likely to be pathogenic. The only observed exonic
change c.739G>T (p.Ala247Ser), was novel, but it was located outside the
functionally important protein domains. Furthermore, this alteration was found at
similar frequencies (ca. 1%) in both cases and controls, suggesting that no
relationship exists between this alteration and a predisposition to cancer. In
addition, none of the six intronic changes seemed to affect consensus splicing
sequences. As similar frequencies were found in cases and controls, none of the
observed changes are likely to be associated with cancer predisposition.
Only one missense change was observed in the 12 exons of AATF, reflecting
extreme conservation of the coding regions of this gene. This preservation could
result from its essential and unique cellular function. Overall, our results provided
no evidence of significant involvement of AATF alterations in familial breast
cancer susceptibility. However, further research is needed to confirm these
findings in other populations and cancer types.
Interestingly, the role of AATF gene in breast cancer has recently been studied
in vitro (Sharma 2013). In that study, the RNomics (small non-mRNAs) of the
AATF gene were analysed in relation to the pathogenesis of breast cancer by
AATF gene silencing and by measuring its effect on the cellular proliferation and
apoptosis induction in human breast cancer cells. The findings suggested that
AATF might be a suitable molecular target for therapeutic interventions for the
cure of cancers. AATF silencing in breast cancer cells can inhibit proliferation,
induce apoptosis, modulate transcription of multiple apoptosis-relevant genes and
71
regulate oestrogen expression in breast cancer cells. (Sharma 2013.) This
highlights the fact that AATF was a good candidate gene, despite the lack of
germline mutations.
6.2
MRG15 and hereditary predisposition to breast cancer (II)
To the best of our knowledge, this study is the first to search for MRG15
mutations in breast cancer patients. The maintenance of genomic integrity and
cell-cycle checkpoint control is dependent on the functioning of MRG15. It
directly binds to PALB2 and this interaction is essential for homologous
recombination regulation during DNA repair. (Smith et al. 2013, Sy et al. 2009a.)
Based on its biological role, it was hypothesized that mutations in the MRG15
gene might contribute to hereditary predisposition to breast cancer.
The whole coding region of this gene was systematically screened for
mutations in breast cancer families. No clearly deleterious sequence alterations
were observed. Thus, no evidence for strong (i.e. high-/moderate-penetrance)
breast cancer predisposing mutations in MRG15 emerged.
In this study, HRM and direct sequencing were used in the mutation search,
and this methodological combination should readily detect all possible deleterious
mutations. The studied cohort consisted of 232 patients derived from both
Northern Finland and Canada. Although no mutations were observed in MRG15
in this study, further studies of larger patient cohorts might still uncover some
very rare disease-related changes in this gene. However, based on our results, it
seems more likely that the MRG15 gene does not make any sizeable contribution
to hereditary breast cancer susceptibility.
Interestingly, after our study was published, Martrat et al. (2011) explored the
link between MRG15 and risk of breast cancer by searching novel components in
the DNA damage repair signalling pathway. They also evaluated the existence of
alterations or mutations of MRG15 in both FA and breast cancer patients, but no
mutations were identified. They pointed out that the role of MRG15 is significant
in the control of genomic stability, but a weak association could not be ruled out
between potential low-penetrance MRG15 variants and breast cancer risk among
BRCA2 mutation carriers. (Martrat et al. 2011.) However, further research is
needed to confirm these findings and to better understand the role of MRG15 in
tumorigenesis.
In both of the two studies discussed above (I and II), no deleterious mutations
in the genes investigated were found. Obviously, the selection of the candidate
72
gene is one of the most determining steps in this kind of targeted mutation
analysis. Most of the candidate genes are chosen solely on the basis of their
function, their association with known cancer predisposition-related pathways, or
their relatedness to other known susceptibility genes, but often the precise
biological functions of these genes remain obscure. However, it is possible that
the function of the additional predisposing genes is unrelated to previously
identified pathways, and therefore, will not be identified through the traditional
candidate gene approach focusing on these.
Equally important is the material studied. While the Finns represent a
genetically isolated population, making disease-related founder mutations easier
to find, the sample size is often limited and geographically may come from a
relatively restricted area.
Candidate gene association studies have been a pioneer in the field of genetic
epidemiology, identifying risk alleles and their associations with various diseases.
The method does, however, have some limitations. There is mild uncertainty
whether the results in general portray disease susceptibility of a common variant
or whether they represent certain ancestral differences existing by chance between
the mixes of test or control populations. A lack of statistical power may also occur
due to small sample size (Patnala et al. 2013). Further, unaffected sisters might
provide a more appropriate comparison group for modern case-control studies
than unrelated population controls, giving more valid relative risk estimates
(Milne et al. 2011).
6.3
PALB2 c.1592delT mutation screening is clinically relevant in
high-risk breast cancer patients tested negative for
BRCA1/BRCA2 (III)
The interpretation of new findings should be applied in clinical genetic services
for the benefit of the majority of women who currently have no explanation for
their breast cancer susceptibility (Southey et al. 2013). As with any test designed
to identify risk factors in the general population, genetic tests for cancer
susceptibility present considerable social and ethical challenges. Many factors
must be taken into account before introducing new genetic markers into clinical
cancer genetics. Careful evaluation is particularly important in the case of alleles
with moderate- or low-penetrance risk. In addition, the limited methods for early
detection and surveillance of certain cancers, e.g. breast cancer vs. pancreatic
tumours, might pose a problem (Foulkes 2008).
73
The current guidelines for deciding which patients should be tested and
which tests performed were updated several years ago. There are differing
opinions about whether, when, and how we should increase the list of genetic
susceptibility markers used in routine clinical care (Ripperger et al. 2009). Recent
clinical recommendations and guidelines already display slight differences across
European countries (Gadzicki et al. 2011).
Biochemical evidence and candidate gene association studies have recently
indicated that BRCA1, BRCA2, and PALB2 are the key breast cancer susceptibility
factors, functioning together in the same important DNA damage response
pathway. PALB2 has emerged as a third major cancer susceptibility factor, as
mutations in this gene have been detected worldwide in most familial breast
cancer cohorts tested. Most pathogenic PALB2 mutations identified to date are
truncating frameshift or nonsense mutations and are found in different parts of the
gene without hot-spot areas (Bogdanova et al. 2011, Erkko et al. 2007, Foulkes et
al. 2007, Garcia et al. 2009b, Hellebrand et al. 2011, Rahman et al. 2007,
Southey et al. 2010) Currently, patients attending clinical genetic services and
seeking advice about their family history of breast cancer are routinely offered
genetic testing for BRCA1 and BRCA2, but not for PALB2 (Southey et al. 2013).
However, based on present knowledge, clinical translation of PALB2 is now
timely.
On average, PALB2 gene mutations confer moderate lifetime risks of breast
cancer, but some mutations in this gene are associated with high breast cancer
risks. These mutations are comparable to the average risk associated with BRCA2
mutations. The cumulative risk of the PALB2 c.1592delT mutation has been
estimated to be 40% (95% CI 17–77) by the age of 70 years, which is almost as
high as for mutations in BRCA2 (45%, 95% CI 31–56) (Erkko et al. 2008,
Southey et al. 2010). In the first study reporting PALB2 c.1592delT in the Finnish
population, the age of cancer onset for mutation carriers appeared to be somewhat
lower than for average population-based breast cancer cases, but still slightly
higher than for BRCA1/2 mutation carriers. However, this finding did not reach
statistical significance (Erkko et al. 2008). In a second study from Southern
Finland, the average age of first breast cancer among c.1592delT carriers was
53.1 years (33.4–79.9), in contrast to 45.2/47.4 years among BRCA1/BRCA2
carriers (Heikkinen et al. 2009). All three groups displayed younger age at disease
onset than the average age of diagnosis in the population in general (61 years)
(National Cancer Institute: SEER 2013). The frequency of different mutations in
PALB2 shows great variability, and ranging from 0.1% to 2.7% worldwide. The
74
Finnish founder mutation has a relatively high frequency, ranging from 0.2% in
healthy controls to 1.8% in unselected breast cancer cases (Erkko et al. 2007),
particularly considering that the 19 most common founder mutations in BRCA1/2
combined account for 1.8% of the unselected cases in Finland (Syrjäkoski et al.
2000). The PALB2 mutation frequency of 4.8% observed in our current clinicalbased study was slightly higher than that reported in the previous study, where
c.1592delT was found in 2.6% of familial breast cancer cases with a strong family
history of the disease (Erkko et al. 2007), but this could also be due to the
relatively small number of families eligible for the analysis.
For women carrying a susceptibility factor and their relatives who might also
be mutation carriers, knowing their mutation status is clinically and
prognostically important. Identified mutation carriers should be kept informed
about optimal breast cancer screening, prevention, or treatment strategies that
could provide a clinical benefit for them. Our data, together with that of other
corresponding studies (Southey et al. 2013), advocate the use of PALB2 gene
mutational testing in non-BRCA families in both diagnostic and predictive testing.
A strength of Study III was that it focused on one well-known mutation. We
know that the c.1592delT mutation causes genomic instability in its carriers,
demonstrated by a significant increase in the total number of spontaneous
chromosomal rearrangements. The significance of this mutation for the
impairment of genomic instability is supported by the results from its recent
molecular and functional assessment. (Nikkilä et al. 2013.)
Personalized medicine for women carrying PALB2 germline mutations might
also be feasible in the short term (Southey et al. 2013). In PALB2 c.1592delT
mutation carriers, PALB2 acts as a haploinsufficient tumour suppressor. LOH
might not always need to be a prerequisite for malignant transformation since the
c.1592delT truncation allele is apparently already on its own sufficient for such
development. This is highly valuable information for developing targeted
therapeutic approaches. (Nikkilä et al. 2013.)
PALB2-deficient cells have been shown to be sensitive to PARP inhibitors,
which selectively target cells deficient in homologous recombination DNA repair,
but only when both alleles are missing or dysfunctional (Buisson et al. 2010).
PARP inhibitors seem not to be suitable for those PALB2 mutation carriers whose
tumours are negative for LOH (Tuma 2009). On the other hand, a patient with
advanced pancreatic cancer revealed biallelic inactivation of the PALB2 gene in
the tumour tissue and received customized treatment with mitomycin C. This
gave a dramatic response and a remarkable clinical outcome (Villarroel et al.
75
2011). Together these findings demonstrate the potential utility of specific drugs
in PALB2 mutation carriers as well as the clinical relevance of providing tailored
therapy for improving outcomes. Testing for this mutation in high-risk breast
cancer patients and in carriers’ siblings could influence the clinical outcome in
many ways, e.g. earlier detection of the malignancy in the context of tailored
surveillance programmes.
The aim of this study was to translate the discoveries of basic science into
clinical practice, thus improving patient care. It is important to note that findings
of studies on cancer susceptibility have a great impact on patients and their
families, including future offspring. A detailed informed consent is therefore
crucial.
6.4
Fanconi anaemia cases in Finland and breast cancer
predisposition (IV)
The link between Fanconi anaemia and hereditary susceptibility to breast cancer
has made all FA genes particularly attractive candidates for cancer-predisposing
genes. Involvement of upstream FA genes in breast cancer susceptibility is,
however, questionable because the link has thus far been established only for
certain homologous recombination-related downstream factors (Howlett et al.
2002, Levran et al. 2005, Reid et al. 2007, Vaz et al. 2010, Xia et al. 2007). This
pathway has prompted several mutational screening studies aimed at identifying
additional breast cancer-predisposing FA genes. The rarity of the FA syndrome
has complicated any conclusive analysis, as natural selection maintains a low
prevalence of the clearly deleterious FA alleles.
This study reports the first mutation confirmed FA patients in Finland,
altogether five patients. All of the identified complementation groups, FA-A, FAG, and FA-I, have been described previously. FANCA is the most commonly
mutated FA gene also in Finland. All FA genes identified encode proteins that are
part of either the core complex (A, G) or the D2-I complex (I), but not part of the
downstream branch in the FA pathway. Three of the FA patients (FA-G, FA-I, and
one FA-A patient) manifested typical FA symptoms, whereas two brothers with
FA-A showed somewhat milder phenotypes. These brothers were both
homozygous for the FANCA c.3384-1G>A splicing mutation, leading to in-frame
deletion of 20 amino acids. Some speculation centres around the defective protein
retaining partial functionality, thus explaining the milder phenotype, but this has
not been tested in functional studies. Curiously, this same mutation has recently
76
been reported in a compound heterozygous state in another FA patient outside
Finland (Chandrasekharappa et al. 2013).
To evaluate the association of identified FA mutations with cancer
susceptibility, their prevalence in breast and prostate cancer cohorts and in
gender-matched controls was tested. All seemed to be recurrent alleles in the
Finnish population and their prevalence did not significantly differ between
cancer cases and controls. In the case of FANCAex1-12del, the family history of
both breast and pancreatic cancer in two of the mutation-positive breast cancer
cases is suggestive of an effect on cancer predisposition in these families, but no
firm evidence for this could be shown. Furthermore, whereas no clearly
pathogenic mutations in the majority of the FA core complex or D2-I complex
have been identified in familial breast cancer re-sequencing efforts (Garcia et al.
2009a, Garcia et al. 2009b, Seal et al. 2003), we did observe deleterious alleles in
affected cases, but they occured at similar frequencies also in unaffected
individuals. This indicates that the tested, clearly deleterious, heterozygous
mutations in the FANCA, FANCG, and FANCI genes do not act as high- or
moderate-risk alleles for breast or prostate cancer. However, since these alleles
seem to be very rare, some low-penetrance effects cannot be excluded. Our results
further strengthen the exclusive role of FA/BRCA pathway downstream genes in
cancer predisposition.
Interestingly, even though the Finnish FANCN/PALB2 c.1592delT founder
mutation exhibits a relatively high frequency in our population, no Finnish FA
patients with this mutation were identified. This indicates that PALB2 c.1592delT
in biallelic state leads to a very severe phenotype (Reid et al. 2007, Xia et al.
2007) or might even be embryonically lethal.
Despite the lack of association with cancer susceptibility, this study provides
clinically relevant and intriguing information about the FA complementation
groups and mutations found in Finland. The clinical diagnosis of FA is not always
easy to make, particularly in cases with atypical clinical features (Liu et al. 2013).
Compared to the found prevalence of heterozygous healthy persons the actual
number of diagnosed cases seems quite limited. Maybe FA is underdiagnosed.
According to genereviews only a minority of the FA patients shows any kind of
dysmorphological features (Genereviews). This study revealed that FA-A, which
is the most common FA subtype worldwide, also appears to be the predominant
subtype in Finland. However, this fact should considered with caution, as the
number of diagnosed and genetically tested Finnish FA patients is quite limited.
77
6.5
Finding the missing heritability of breast cancer susceptibility
Over the past few years, the term “missing heritability” has been used to
described the situation that most variants identified thus far confer relatively
small increments in breast cancer risk and explain only a minor proportion of
familial clustering, leading to the question of how the remaining cases can be
explained (Manolio et al. 2009). Although GWA studies have been successful in
identifying common variants involved in complex diseases, such as cancer, for the
majority of complex traits <10% of the genetic variance is explained by common
variants (Eichler et al. 2010). Despite a tremendous amount of work, the
candidate gene approach to uncover new breast cancer susceptibility genes has
offered limited success – even in cases that initially appeared quite promising. At
present, known susceptibility genes are thought to be responsible for less than
one-third of familial breast cancer cases. Although we believe that most of the
known susceptibility genes function in the maintenance of genomic integrity, we
cannot be sure that the missing genes will be related to similar biological
pathways.
Obviously, the missing heritability is a far more complicated issue than is
currently realized and its understanding requires a more comprehensive
assessment. It has been suggested that much of the missing heritability lies in
gene-gene interactions. In a Mendelian disorder, one main gene acting in concert
with other gene variants might be involved (Kaiser 2012). Recently, seven leading
geneticists speculated about where the missing heritability might be found
(Eichler et al. 2010). Perhaps a myriad of common variants of small effect does
explain the vast majority of genetic effects predicted by heritability estimates,
even though we cannot detect them individually. Further, the heritability could be
due to the joint action of numerous loci of small effect. Gene-environment
interactions may be important as well, but some susceptibility variants might
confer risk only when inherited from a specific parent (parent-of-origin effect).
Epigenetic effects beyond imprinting, such as DNA methylation and histone
modifications, are examples of the complex relationship between genotype and
phenotype. The discovery of non-coding microRNAs has revealed a new
mechanism of translational regulation, and SNPs associated with breast cancer
susceptibility have been shown to alter microRNA gene regulation. It is also
reasonable to presume that coding sequence variations could synergistically act
through protein-protein interactions and protein-DNA interactions in
transcriptional networks and biochemical systems. One further hypothesis is that
78
a significant portion of the missing heritability is due to neither single common
variants nor single rare variants, but rather to rare combinations of common
variants. Finally, the possibility that some heritability comes from entirely
unforeseen sources is something to look forward to exploring in the future.
(Eichler et al. 2010.)
79
80
7
Concluding remarks and future prospects
The main purpose of this thesis was to elucidate the background of familial breast
cancer susceptibility. Currently, only ~30% of hereditary breast cancers can be
explained by mutations in known breast cancer susceptibility genes. The detailed
action mechanisms of these mutations and their exact penetrance remain largely
obscure. New familial breast cancer predisposing factors need to be identified
alongside further evaluation of those already recognized. When sufficiently
validated, the knowledge gained in research laboratories should be transferred
into clinical practice. Resolving the factors contributing to tumorigenesis, such as
uncovering new predisposing alleles, will ultimately lead to a better
understanding of cancer development, disease progression, and treatment. The
putative candidate genes chosen for this study were AATF and MRG15. They
were chosen because of their biological functions in the maintenance of genomic
integrity and/or their interactions with known susceptibility gene products. The
PALB2 c.1592delT mutation was evaluated as a clinical tool for diagnostic
aetiology in high-risk breast cancer families. The DNA samples of known Finnish
FA patients were studied in order to identify their complementation groups and
the causative mutations. Additional evaluation was carried out to determine
whether these FA gene mutations are enriched in breast and prostate cancer
patient series. Based on the results of each study, the following observations and
conclusions were drawn:
1.
2.
3.
4.
No clearly pathogenic mutations were found in the AATF gene. The
contribution of AATF germline mutations to breast cancer predisposition thus
appears rare or absent.
Germline mutations in the MRG15 gene do not make any sizeable
contributions to familial breast cancer, at least not in the populations studied.
The PALB2 c.1592delT founder mutation is known to be associated with a
six-fold increased breast cancer risk. Based on our results, PALB2 c.1592delT
should be included in the routine scheme of gene tests in genetic counselling,
particularly for BRCA1/2-negative high-risk breast cancer patients. Mutation
carriers should have tailored surveillance programmes.
The complementation groups and gene mutations of Finnish Fanconi anaemia
patients were identified, but no relationship between FA mutations and breast
or prostate cancer susceptibility was observed, suggesting that deleterious
81
FANCA, FANCG, and FANCI mutations are not high- or moderate-risk alleles
for breast or prostate cancer.
To summarize, this study shed light on the background of familial breast cancer
and forged some novel paths in a clinical setting. However, more research is
needed to better understand tumorigenesis and to uncover all genetic factors as
well as other factors predisposing to breast cancer. A considerable proportion of
the remaining familial breast and/or ovarian cancer cases could be due to the
multiplicative effect of various low- to moderate-penetrance genes, in
combination with environmental and lifestyle factors. Since the genetic
contribution to breast cancer development and the whole process of tumorigenesis
remain largely unknown, extensive investigations in these areas will continue.
Current studies may be focusing too much on monogenic alterations; because
cancer is a polygenic disease, researchers should perhaps widen their scope to
gene-gene interactions, gene-environment interactions, and epigenetic changes.
Better understanding of the whole process leading to breast cancer and defining
the specific molecular defects involved may stymie disease development and,
ultimately, bring benefits in the fight against breast cancer.
82
References
Prevalence and penetrance of BRCA1 and BRCA2 mutations in a population-based series
of breast cancer cases. Anglian Breast Cancer Study Group. (2000) Br J Cancer
83(10): 1301–1308.
Aaltonen LA, Peltomäki P, Leach FS, Sistonen P, Pylkkänen L, Mecklin JP, Järvinen H,
Powell SM, Jen J, Hamilton SR, Petersen GM, Kinzler KW, Vogelstein B & de la
Chapelle A (1993) Clues to the pathogenesis of familial colorectal cancer. Science
260: 812–816.
Abraham RT (2004) PI 3-kinase related kinases: 'big' players in stress-induced signaling
pathways. DNA Repair (Amst) 3(8–9): 883–887.
Ahmed M & Rahman N (2006) ATM and breast cancer susceptibility. Oncogene 25(43):
5906–5911.
Akbari MR, Ghadirian P, Robidoux A, Foumani M, Sun Y, Royer R, Zandvakili I, Lynch
H & Narod SA (2009) Germline RAP80 mutations and susceptibility to breast cancer.
Breast Cancer Res Treat 113(2): 377–381.
Akbari MR, Tonin P, Foulkes WD, Ghadirian P, Tischkowitz M & Narod SA (2010)
RAD51C germline mutations in breast and ovarian cancer patients. Breast Cancer Res
12(4): 404.
Alexandrov LB, Nik-Zainal S, Wedge DC, Aparicio SA, Behjati S, Biankin AV, Bignell
GR, Bolli N, Borg A, Borresen-Dale AL, Boyault S, Burkhardt B, Butler AP, Caldas
C, Davies HR, Desmedt C, Eils R, Eyfjord JE, Foekens JA, Greaves M, Hosoda F,
Hutter B, Ilicic T, Imbeaud S, Imielinsk M, Jager N, Jones DT, Jones D, Knappskog S,
Kool M, Lakhani SR, Lopez-Otin C, Martin S, Munshi NC, Nakamura H, Northcott
PA, Pajic M, Papaemmanuil E, Paradiso A, Pearson JV, Puente XS, Raine K,
Ramakrishna M, Richardson AL, Richter J, Rosenstiel P, Schlesner M, Schumacher
TN, Span PN, Teague JW, Totoki Y, Tutt AN, Valdes-Mas R, van Buuren MM, van 't
Veer L, Vincent-Salomon A, Waddell N, Yates LR, Australian Pancreatic Cancer
Genome Initiative, ICGC Breast Cancer Consortium, ICGC MMML-Seq Consortium,
ICGC PedBrain, Zucman-Rossi J, Futreal PA, McDermott U, Lichter P, Meyerson M,
Grimmond SM, Siebert R, Campo E, Shibata T, Pfister SM, Campbell PJ & Stratton
MR (2013) Signatures of mutational processes in human cancer. Nature 500(7463):
415–421.
Allinen M, Huusko P, Mäntyniemi S, Launonen V & Winqvist R (2001) Mutation analysis
of the CHK2 gene in families with hereditary breast cancer. Br J Cancer 85(2): 209–
212.
83
Antoniou A, Pharoah PD, Narod S, Risch HA, Eyfjord JE, Hopper JL, Loman N, Olsson H,
Johannsson O, Borg A, Pasini B, Radice P, Manoukian S, Eccles DM, Tang N, Olah E,
Anton-Culver H, Warner E, Lubinski J, Gronwald J, Gorski B, Tulinius H, Thorlacius
S, Eerola H, Nevanlinna H, Syrjäkoski K, Kallioniemi OP, Thompson D, Evans C,
Peto J, Lalloo F, Evans DG & Easton DF (2003) Average risks of breast and ovarian
cancer associated with BRCA1 or BRCA2 mutations detected in case Series
unselected for family history: a combined analysis of 22 studies. Am J Hum Genet
72(5): 1117–1130.
Antoniou AC & Easton DF (2003) Polygenic inheritance of breast cancer: Implications for
design of association studies. Genet Epidemiol 25(3): 190–202.
Antoniou AC & Easton DF (2006) Models of genetic susceptibility to breast cancer.
Oncogene 25(43): 5898–5905.
Apostolou P & Fostira F (2013) Hereditary breast cancer: the era of new susceptibility
genes. Biomed Res Int 2013: 747318.
Araten DJ, Golde DW, Zhang RH, Thaler HT, Gargiulo L, Notaro R & Luzzatto L (2005)
A quantitative measurement of the human somatic mutation rate. Cancer Res 65(18):
8111–8117.
Barrett SV (2010) Breast cancer. J R Coll Physicians Edinb 40(4): 335–8; quiz 339.
Bartkova J, Tommiska J, Oplustilova L, Aaltonen K, Tamminen A, Heikkinen T, Mistrik
M, Aittomäki K, Blomqvist C, Heikkilä P, Lukas J, Nevanlinna H & Bartek J (2008)
Aberrations of the MRE11-RAD50-NBS1 DNA damage sensor complex in human
breast cancer: MRE11 as a candidate familial cancer-predisposing gene. Mol Oncol
2(4): 296–316.
Beggs AD, Latchford AR, Vasen HF, Moslein G, Alonso A, Aretz S, Bertario L, Blanco I,
Bulow S, Burn J, Capella G, Colas C, Friedl W, Moller P, Hes FJ, Jarvinen H,
Mecklin JP, Nagengast FM, Parc Y, Phillips RK, Hyer W, Ponz de Leon M,
Renkonen-Sinisalo L, Sampson JR, Stormorken A, Tejpar S, Thomas HJ, Wijnen JT,
Clark SK & Hodgson SV (2010) Peutz-Jeghers syndrome: a systematic review and
recommendations for management. Gut 59(7): 975–986.
Bekker-Jensen S & Mailand N (2010) Assembly and function of DNA double-strand break
repair foci in mammalian cells. DNA Repair (Amst) 9(12): 1219–1228.
Berg JW & Hutter RV (1995) Breast cancer. Cancer 75(1 Suppl): 257–269.
Bertram MJ & Pereira-Smith OM (2001) Conservation of the MORF4 related gene family:
identification of a new chromo domain subfamily and novel protein motif. Gene
266(1–2): 111–121.
Bishop JM (1981) Enemies within: the genesis of retrovirus oncogenes. Cell 23(1): 5–6.
Bishop JM (1991) Molecular themes in oncogenesis. Cell 64(2): 235–248.
Bleuyard JY, Buisson R, Masson JY & Esashi F (2012) ChAM, a novel motif that
mediates PALB2 intrinsic chromatin binding and facilitates DNA repair. EMBO Rep
13(2): 135–141.
Blom E, van de Vrugt HJ, de Vries Y, de Winter JP, Arwert F & Joenje H (2004) Multiple
TPR motifs characterize the Fanconi anemia FANCG protein. DNA Repair (Amst)
3(1): 77–84.
84
BLOOM HJ & RICHARDSON WW (1957) Histological grading and prognosis in breast
cancer; a study of 1409 cases of which 359 have been followed for 15 years. Br J
Cancer 11(3): 359–377.
Bogdanova N, Feshchenko S, Schurmann P, Waltes R, Wieland B, Hillemanns P, Rogov
YI, Dammann O, Bremer M, Karstens JH, Sohn C, Varon R & Dork T (2008)
Nijmegen Breakage Syndrome mutations and risk of breast cancer. Int J Cancer
122(4): 802–806.
Bogdanova N, Sokolenko AP, Iyevleva AG, Abysheva SN, Blaut M, Bremer M,
Christiansen H, Rave-Frank M, Dork T & Imyanitov EN (2011) PALB2 mutations in
German and Russian patients with bilateral breast cancer. Breast Cancer Res Treat
126(2): 545–550.
Bogliolo M, Schuster B, Stoepker C, Derkunt B, Su Y, Raams A, Trujillo JP, Minguillon J,
Ramirez MJ, Pujol R, Casado JA, Banos R, Rio P, Knies K, Zuniga S, Benitez J,
Bueren JA, Jaspers NG, Scharer OD, de Winter JP, Schindler D & Surralles J (2013)
Mutations in ERCC4, encoding the DNA-repair endonuclease XPF, cause Fanconi
anemia. Am J Hum Genet 92(5): 800–806.
Boulton SJ (2006) BRCA1-mediated ubiquitylation. Cell Cycle 5(14): 1481–1486.
Bradbury AR & Olopade OI (2007) Genetic susceptibility to breast cancer. Rev Endocr
Metab Disord 8(3): 255–267.
Broca PP (1866) Traité des Tumeurs1. Paris, P. Asselin.
Brownstein MH, Wolf M & Bikowski JB (1978) Cowden's disease: a cutaneous marker of
breast cancer. Cancer 41(6): 2393–2398.
Brunak S, Engelbrecht J & Knudsen S (1991) Prediction of human mRNA donor and
acceptor sites from the DNA sequence. J Mol Biol 220(1): 49–65.
Bruno T, De Angelis R, De Nicola F, Barbato C, Di Padova M, Corbi N, Libri V, Benassi
B, Mattei E, Chersi A, Soddu S, Floridi A, Passananti C & Fanciulli M (2002) Che-1
affects cell growth by interfering with the recruitment of HDAC1 by Rb. Cancer Cell
2(5): 387–399.
Bruno T, De Nicola F, Iezzi S, Lecis D, D'Angelo C, Di Padova M, Corbi N, Dimiziani L,
Zannini L, Jekimovs C, Scarsella M, Porrello A, Chersi A, Crescenzi M, Leonetti C,
Khanna KK, Soddu S, Floridi A, Passananti C, Delia D & Fanciulli M (2006) Che-1
phosphorylation by ATM/ATR and Chk2 kinases activates p53 transcription and the
G2/M checkpoint. Cancer Cell 10(6): 473–486.
Bruno T, Desantis A, Bossi G, Di Agostino S, Sorino C, De Nicola F, Iezzi S, Franchitto A,
Benassi B, Galanti S, La Rosa F, Floridi A, Bellacosa A, Passananti C, Blandino G &
Fanciulli M (2010) Che-1 promotes tumor cell survival by sustaining mutant p53
transcription and inhibiting DNA damage response activation. Cancer Cell 18(2):
122–134.
Bryant HE, Schultz N, Thomas HD, Parker KM, Flower D, Lopez E, Kyle S, Meuth M,
Curtin NJ & Helleday T (2005) Specific killing of BRCA2-deficient tumours with
inhibitors of poly(ADP-ribose) polymerase. Nature 434(7035): 913–917.
85
Buisson R, Dion-Cote AM, Coulombe Y, Launay H, Cai H, Stasiak AZ, Stasiak A, Xia B
& Masson JY (2010) Cooperation of breast cancer proteins PALB2 and piccolo
BRCA2 in stimulating homologous recombination. Nat Struct Mol Biol 17(10): 1247–
1254.
Bunz F (2008) Principles of Cancer Genetics. Baltimore, USA, Springer.
Burge C & Karlin S (1997) Prediction of complete gene structures in human genomic
DNA. J Mol Biol 268(1): 78–94.
Byrnes GB, Southey MC & Hopper JL (2008) Are the so-called low penetrance breast
cancer genes, ATM, BRIP1, PALB2 and CHEK2, high risk for women with strong
family histories? Breast Cancer Res 10(3): 208.
Caestecker KW & Van de Walle GR (2013) The role of BRCA1 in DNA double-strand
repair: past and present. Exp Cell Res 319(5): 575–587.
Campos B, Balmana J, Gardenyes J, Valenzuela I, Abad O, Fabregas P, Volpini V & Diez
O (2013) Germline mutations in NF1 and BRCA1 in a family with neurofibromatosis
type 1 and early-onset breast cancer. Breast Cancer Res Treat 139(2): 597–602.
Cantor SB, Bell DW, Ganesan S, Kass EM, Drapkin R, Grossman S, Wahrer DC, Sgroi
DC, Lane WS, Haber DA & Livingston DM (2001) BACH1, a novel helicase-like
protein, interacts directly with BRCA1 and contributes to its DNA repair function.
Cell 105(1): 149–160.
Cao AY, Huang J, Hu Z, Li WF, Ma ZL, Tang LL, Zhang B, Su FX, Zhou J, Di GH, Shen
KW, Wu J, Lu JS, Luo JM, Yuan WT, Shen ZZ, Huang W & Shao ZM (2009) The
prevalence of PALB2 germline mutations in BRCA1/BRCA2 negative Chinese
women with early onset breast cancer or affected relatives. Breast Cancer Res Treat
114(3): 457–462.
Carrozza MJ, Li B, Florens L, Suganuma T, Swanson SK, Lee KK, Shia WJ, Anderson S,
Yates J, Washburn MP & Workman JL (2005) Histone H3 methylation by Set2
directs deacetylation of coding regions by Rpd3S to suppress spurious intragenic
transcription. Cell 123(4): 581–592.
Cartegni L, Wang J, Zhu Z, Zhang MQ & Krainer AR (2003) ESEfinder: A web resource
to identify exonic splicing enhancers. Nucleic Acids Res 31(13): 3568–3571.
Chandra S, Levran O, Jurickova I, Maas C, Kapur R, Schindler D, Henry R, Milton K,
Batish SD, Cancelas JA, Hanenberg H, Auerbach AD & Williams DA (2005) A rapid
method for retrovirus-mediated identification of complementation groups in Fanconi
anemia patients. Mol Ther 12(5): 976–984.
Chandrasekharappa SC, Lach FP, Kimble DC, Kamat A, Teer JK, Donovan FX, Flynn E,
Sen SK, Thongthip S, Sanborn E, Smogorzewska A, Auerbach AD, Ostrander EA &
NISC Comparative Sequencing Program (2013) Massively parallel sequencing, aCGH,
and RNA-Seq technologies provide a comprehensive molecular diagnosis of Fanconi
anemia. Blood 121(22): e138–48.
CHEK2 Breast Cancer Case-Control Consortium (2004) CHEK2*1100delC and
susceptibility to breast cancer: a collaborative analysis involving 10,860 breast cancer
cases and 9,065 controls from 10 studies. Am J Hum Genet 74(6): 1175–1182.
86
Chen M, Tominaga K & Pereira-Smith OM (2010) Emerging role of the MORF/MRG
gene family in various biological processes, including aging. Ann N Y Acad Sci 1197:
134–141.
Chenevix-Trench G, Spurdle AB, Gatei M, Kelly H, Marsh A, Chen X, Donn K,
Cummings M, Nyholt D, Jenkins MA, Scott C, Pupo GM, Dork T, Bendix R, Kirk J,
Tucker K, McCredie MR, Hopper JL, Sambrook J, Mann GJ & Khanna KK (2002)
Dominant negative ATM mutations in breast cancer families. J Natl Cancer Inst 94(3):
205–215.
Cheng NC, van de Vrugt HJ, van der Valk MA, Oostra AB, Krimpenfort P, de Vries Y,
Joenje H, Berns A & Arwert F (2000) Mice with a targeted disruption of the Fanconi
anemia homolog Fanca. Hum Mol Genet 9(12): 1805–1811.
Chompret A, Brugieres L, Ronsin M, Gardes M, Dessarps-Freichey F, Abel A, Hua D,
Ligot L, Dondon MG, Bressac-de Paillerets B, Frebourg T, Lemerle J, Bonaiti-Pellie
C & Feunteun J (2000) P53 germline mutations in childhood cancers and cancer risk
for carrier individuals. Br J Cancer 82(12): 1932–1937.
Chow LM & Baker SJ (2006) PTEN function in normal and neoplastic growth. Cancer
Lett 241(2): 184–196.
Ciccia A & Elledge SJ (2010) The DNA damage response: making it safe to play with
knives. Mol Cell 40(2): 179–204.
Ciciarello M, Mangiacasale R, Casenghi M, Zaira Limongi M, D'Angelo M, Soddu S,
Lavia P & Cundari E (2001) p53 displacement from centrosomes and p53-mediated
G1 arrest following transient inhibition of the mitotic spindle. J Biol Chem 276(22):
19205–19213.
Claus EB, Schildkraut JM, Thompson WD & Risch NJ (1996) The genetic attributable risk
of breast and ovarian cancer. Cancer 77(11): 2318–2324.
Cohn MA, Kowal P, Yang K, Haas W, Huang TT, Gygi SP & D'Andrea AD (2007) A
UAF1-containing multisubunit protein complex regulates the Fanconi anemia
pathway. Mol Cell 28(5): 786–797.
Collaborative Group on Hormonal Factors in Breast Cancer (2001) Familial breast cancer:
collaborative reanalysis of individual data from 52 epidemiological studies including
58,209 women with breast cancer and 101,986 women without the disease. Lancet
358(9291): 1389–1399.
Crossan GP & Patel KJ (2012) The Fanconi anaemia pathway orchestrates incisions at
sites of crosslinked DNA. J Pathol 226(2): 326–337.
D'Andrea AD (2010) Susceptibility pathways in Fanconi's anemia and breast cancer. N
Engl J Med 362(20): 1909–1919.
Davies AA, Masson JY, McIlwraith MJ, Stasiak AZ, Stasiak A, Venkitaraman AR & West
SC (2001) Role of BRCA2 in control of the RAD51 recombination and DNA repair
protein. Mol Cell 7(2): 273–282.
87
De Leeneer K, Van Bockstal M, De Brouwer S, Swietek N, Schietecatte P, Sabbaghian N,
Van den Ende J, Willocx S, Storm K, Blaumeiser B, Van Asperen CJ, Wijnen JT,
Leunen K, Legius E, Michils G, Matthijs G, Blok MJ, Gomez-Garcia E, De Paepe A,
Tischkowitz M, Poppe B & Claes K (2012) Evaluation of RAD51C as cancer
susceptibility gene in a large breast-ovarian cancer patient population referred for
genetic testing. Breast Cancer Res Treat 133(1): 393–398.
De Nicola F, Bruno T, Iezzi S, Di Padova M, Floridi A, Passananti C, Del Sal G &
Fanciulli M (2007) The prolyl isomerase Pin1 affects Che-1 stability in response to
apoptotic DNA damage. J Biol Chem 282(27): 19685–19691.
Devilee P, Cleton-Jansen AM & Cornelisse CJ (2001) Ever since Knudson. Trends Genet
17(10): 569–573.
di Masi A & Antoccia A (2008) NBS1 Heterozygosity and Cancer Risk. Curr Genomics
9(4): 275–281.
Docrates syöpäsairaala (2013) Geeniprofilointi syövänhoidon kohdentamisen apuna
Docrateella.
URI:
http://www.docrates.com/fi/syovan-hoito/uudet-hoitomuodot/
geeniprofilointi-syovanhoidon-kohdentamisen-apuna-docrateella.
Domchek SM, Tang J, Stopfer J, Lilli DR, Hamel N, Tischkowitz M, Monteiro AN,
Messick TE, Powers J, Yonker A, Couch FJ, Goldgar DE, Davidson HR, Nathanson
KL, Foulkes WD & Greenberg RA (2013) Biallelic deleterious BRCA1 mutations in a
woman with early-onset ovarian cancer. Cancer Discov 3(4): 399–405.
Dorsman JC, Levitus M, Rockx D, Rooimans MA, Oostra AB, Haitjema A, Bakker ST,
Steltenpool J, Schuler D, Mohan S, Schindler D, Arwert F, Pals G, Mathew CG,
Waisfisz Q, de Winter JP & Joenje H (2007) Identification of the Fanconi anemia
complementation group I gene, FANCI. Cell Oncol 29(3): 211–218.
Downs JA, Nussenzweig MC & Nussenzweig A (2007) Chromatin dynamics and the
preservation of genetic information. Nature 447(7147): 951–958.
Downs-Holmes C & Silverman P (2011) Breast cancer: overview & updates. Nurse Pract
36(12): 20–6; quiz 7.
Doyon Y, Selleck W, Lane WS, Tan S & Cote J (2004) Structural and functional
conservation of the NuA4 histone acetyltransferase complex from yeast to humans.
Mol Cell Biol 24(5): 1884–1896.
Dray E, Etchin J, Wiese C, Saro D, Williams GJ, Hammel M, Yu X, Galkin VE, Liu D,
Tsai MS, Sy SM, Schild D, Egelman E, Chen J & Sung P (2010) Enhancement of
RAD51 recombinase activity by the tumor suppressor PALB2. Nat Struct Mol Biol
17(10): 1255–1259.
Dzikiewicz-Krawczyk A (2008) The importance of making ends meet: mutations in genes
and altered expression of proteins of the MRN complex and cancer. Mutat Res 659(3):
262–273.
Easton DF (1994) Cancer risks in A-T heterozygotes. Int J Radiat Biol 66(6 Suppl): S177–
82.
88
Easton DF, Pooley KA, Dunning AM, Pharoah PD, Thompson D, Ballinger DG,
Struewing JP, Morrison J, Field H, Luben R, Wareham N, Ahmed S, Healey CS,
Bowman R, SEARCH collaborators, Meyer KB, Haiman CA, Kolonel LK, Henderson
BE, Le Marchand L, Brennan P, Sangrajrang S, Gaborieau V, Odefrey F, Shen CY,
Wu PE, Wang HC, Eccles D, Evans DG, Peto J, Fletcher O, Johnson N, Seal S,
Stratton MR, Rahman N, Chenevix-Trench G, Bojesen SE, Nordestgaard BG,
Axelsson CK, Garcia-Closas M, Brinton L, Chanock S, Lissowska J, Peplonska B,
Nevanlinna H, Fagerholm R, Eerola H, Kang D, Yoo KY, Noh DY, Ahn SH, Hunter
DJ, Hankinson SE, Cox DG, Hall P, Wedren S, Liu J, Low YL, Bogdanova N,
Schurmann P, Dork T, Tollenaar RA, Jacobi CE, Devilee P, Klijn JG, Sigurdson AJ,
Doody MM, Alexander BH, Zhang J, Cox A, Brock IW, MacPherson G, Reed MW,
Couch FJ, Goode EL, Olson JE, Meijers-Heijboer H, van den Ouweland A,
Uitterlinden A, Rivadeneira F, Milne RL, Ribas G, Gonzalez-Neira A, Benitez J,
Hopper JL, McCredie M, Southey M, Giles GG, Schroen C, Justenhoven C, Brauch H,
Hamann U, Ko YD, Spurdle AB, Beesley J, Chen X, kConFab, AOCS Management
Group, Mannermaa A, Kosma VM, Kataja V, Hartikainen J, Day NE, Cox DR &
Ponder BA (2007) Genome-wide association study identifies novel breast cancer
susceptibility loci. Nature 447(7148): 1087–1093.
Eichler EE, Flint J, Gibson G, Kong A, Leal SM, Moore JH & Nadeau JH (2010) Missing
heritability and strategies for finding the underlying causes of complex disease. Nat
Rev Genet 11(6): 446–450.
Elledge SJ (1996) Cell cycle checkpoints: preventing an identity crisis. Science 274(5293):
1664–1672.
Ellsworth RE, Decewicz DJ, Shriver CD & Ellsworth DL (2010) Breast cancer in the
personal genomics era. Curr Genomics 11(3): 146–161.
Elston CW & Ellis IO (1991) Pathological prognostic factors in breast cancer. I. The value
of histological grade in breast cancer: experience from a large study with long-term
follow-up. Histopathology 19(5): 403–410.
Erkko H, Dowty JG, Nikkilä J, Syrjäkoski K, Mannermaa A, Pylkäs K, Southey MC, Holli
K, Kallioniemi A, Jukkola-Vuorinen A, Kataja V, Kosma VM, Xia B, Livingston DM,
Winqvist R & Hopper JL (2008) Penetrance analysis of the PALB2 c.1592delT
founder mutation. Clin Cancer Res 14(14): 4667–4671.
Erkko H, Xia B, Nikkilä J, Schleutker J, Syrjäkoski K, Mannermaa A, Kallioniemi A,
Pylkäs K, Karppinen SM, Rapakko K, Miron A, Sheng Q, Li G, Mattila H, Bell DW,
Haber DA, Grip M, Reiman M, Jukkola-Vuorinen A, Mustonen A, Kere J, Aaltonen
LA, Kosma VM, Kataja V, Soini Y, Drapkin RI, Livingston DM & Winqvist R (2007)
A recurrent mutation in PALB2 in Finnish cancer families. Nature 446(7133): 316–
319.
Esashi F, Christ N, Gannon J, Liu Y, Hunt T, Jasin M & West SC (2005) CDK-dependent
phosphorylation of BRCA2 as a regulatory mechanism for recombinational repair.
Nature 434(7033): 598–604.
Evers B & Jonkers J (2006) Mouse models of BRCA1 and BRCA2 deficiency: past
lessons, current understanding and future prospects. Oncogene 25(43): 5885–5897.
89
Fanciulli M, Bruno T, Di Padova M, De Angelis R, Iezzi S, Iacobini C, Floridi A &
Passananti C (2000) Identification of a novel partner of RNA polymerase II subunit 11,
Che-1, which interacts with and affects the growth suppression function of Rb.
FASEB J 14(7): 904–912.
Fanconi anaemia/Breast cancer consortium (1996) Positional cloning of the Fanconi
anaemia group A gene. Nat Genet 14(3): 324–328.
Farmer H, McCabe N, Lord CJ, Tutt AN, Johnson DA, Richardson TB, Santarosa M,
Dillon KJ, Hickson I, Knights C, Martin NM, Jackson SP, Smith GC & Ashworth A
(2005) Targeting the DNA repair defect in BRCA mutant cells as a therapeutic
strategy. Nature 434(7035): 917–921.
Fearon ER & Vogelstein B (1990) A genetic model for colorectal tumorigenesis. Cell
61(5): 759–767.
Ferlay J, Steliarova-Foucher E, Lortet-Tieulent J, Rosso S, Coebergh JW, Comber H,
Forman D & Bray F (2013) Cancer incidence and mortality patterns in Europe:
estimates for 40 countries in 2012. Eur J Cancer 49(6): 1374–1403.
Finnish Cancer Registry NORDCAN. URI: http://www-dep.iarc.fr/NORDCAN/FI/
StatsFact.asp?cancer=180&country=246.
Fishel R & Kolodner RD (1995) Identification of mismatch repair genes and their role in
the development of cancer. Curr Opin Genet Dev 5(3): 382–395.
Folsom AR, Kaye SA, Prineas RJ, Potter JD, Gapstur SM & Wallace RB (1990) Increased
incidence of carcinoma of the breast associated with abdominal adiposity in
postmenopausal women. Am J Epidemiol 131(5): 794–803.
Fong PC, Boss DS, Yap TA, Tutt A, Wu P, Mergui-Roelvink M, Mortimer P, Swaisland H,
Lau A, O'Connor MJ, Ashworth A, Carmichael J, Kaye SB, Schellens JH & de Bono
JS (2009) Inhibition of poly(ADP-ribose) polymerase in tumors from BRCA mutation
carriers. N Engl J Med 361(2): 123–134.
Foulkes WD (2008) Inherited susceptibility to common cancers. N Engl J Med 359(20):
2143–2153.
Foulkes WD, Ghadirian P, Akbari MR, Hamel N, Giroux S, Sabbaghian N, Darnel A,
Royer R, Poll A, Fafard E, Robidoux A, Martin G, Bismar TA, Tischkowitz M,
Rousseau F & Narod SA (2007) Identification of a novel truncating PALB2 mutation
and analysis of its contribution to early-onset breast cancer in French-Canadian
women. Breast Cancer Res 9(6): R83.
Francken AB, Schouten PC, Bleiker EM, Linn SC & Rutgers EJ (2013) Breast cancer in
women at high risk: the role of rapid genetic testing for BRCA1 and -2 mutations and
the consequences for treatment strategies. Breast 22(5): 561–568.
Frank TS (2001) Hereditary cancer syndromes. Arch Pathol Lab Med 125(1): 85–90.
Gadzicki D, Evans DG, Harris H, Julian-Reynier C, Nippert I, Schmidtke J, Tibben A, van
Asperen CJ & Schlegelberger B (2011) Genetic testing for familial/hereditary breast
cancer-comparison of guidelines and recommendations from the UK, France, the
Netherlands and Germany. J Community Genet 2(2): 53–69.
90
Ganguly A, Rock MJ & Prockop DJ (1993) Conformation-sensitive gel electrophoresis for
rapid detection of single-base differences in double-stranded PCR products and DNA
fragments: evidence for solvent-induced bends in DNA heteroduplexes. Proc Natl
Acad Sci USA 90(21): 10325–10329.
Garcia MJ, Fernandez V, Osorio A, Barroso A, Fernandez F, Urioste M & Benitez J
(2009a) Mutational analysis of FANCL, FANCM and the recently identified FANCI
suggests that among the 13 known Fanconi Anemia genes, only FANCD1/BRCA2
plays a major role in high-risk breast cancer predisposition. Carcinogenesis 30(11):
1898–1902.
Garcia MJ, Fernandez V, Osorio A, Barroso A, Llort G, Lazaro C, Blanco I, Caldes T, de
la Hoya M, Ramon Y Cajal T, Alonso C, Tejada MI, San Roman C, Robles-Diaz L,
Urioste M & Benitez J (2009b) Analysis of FANCB and FANCN/PALB2 fanconi
anemia genes in BRCA1/2-negative Spanish breast cancer families. Breast Cancer Res
Treat 113(3): 545–551.
Garcia SN, Kirtane BM, Podlutsky AJ, Pereira-Smith OM & Tominaga K (2007) Mrg15
null and heterozygous mouse embryonic fibroblasts exhibit DNA-repair defects post
exposure to gamma ionizing radiation. FEBS Lett 581(27): 5275–5281.
Garcia-Higuera I, Taniguchi T, Ganesan S, Meyn MS, Timmers C, Hejna J, Grompe M &
D'Andrea AD (2001) Interaction of the Fanconi anemia proteins and BRCA1 in a
common pathway. Mol Cell 7(2): 249–262.
Geary J, Sasieni P, Houlston R, Izatt L, Eeles R, Payne SJ, Fisher S & Hodgson SV (2008)
Gene-related cancer spectrum in families with hereditary non-polyposis colorectal
cancer (HNPCC). Fam Cancer 7(2): 163–172.
Genereviews (2013) Fanconi Anemia. URI: www.ncbi.nlm.nih.gov/books/NBK1401/
German J, Schonberg S, Caskie S, Warburton D, Falk C & Ray JH (1987) A test for
Fanconi's anemia. Blood 69(6): 1637–1641.
Gervaz P, Cerottini JP, Bouzourene H, Hahnloser D, Doan CL, Benhattar J, Chaubert P,
Secic M, Gillet M & Carethers JM (2002) Comparison of microsatellite instability and
chromosomal instability in predicting survival of patients with T3N0 colorectal cancer.
Surgery 131(2): 190–197.
GLOBOCAN 2008 WHO, International Agengy for Research on Cancer. URI:
http://globocan.iarc.fr/factsheet.asp.
Goldstein DB (2009) Common genetic variation and human traits. N Engl J Med 360(17):
1696–1698.
Guilford P, Hopkins J, Harraway J, McLeod M, McLeod N, Harawira P, Taite H, Scoular
R, Miller A & Reeve AE (1998) E-cadherin germline mutations in familial gastric
cancer. Nature 392(6674): 402–405.
Guilford P, Humar B & Blair V (2010) Hereditary diffuse gastric cancer: translation of
CDH1 germline mutations into clinical practice. Gastric Cancer 13(1): 1–10.
Guo Q & Xie J (2004) AATF inhibits aberrant production of amyloid beta peptide 1–42 by
interacting directly with Par-4. J Biol Chem 279(6): 4596–4603.
91
Haanpää M, Pylkäs K, Moilanen JS & Winqvist R (2013) Evaluation of the need for
routine clinical testing of PALB2 c.1592delT mutation in BRCA negative Northern
Finnish breast cancer families. BMC Med Genet 14(1): 82.
Hahn WC, Counter CM, Lundberg AS, Beijersbergen RL, Brooks MW & Weinberg RA
(1999) Creation of human tumour cells with defined genetic elements. Nature
400(6743): 464–468.
Hall JM, Lee MK, Newman B, Morrow JE, Anderson LA, Huey B & King MC (1990)
Linkage of early-onset familial breast cancer to chromosome 17q21. Science
250(4988): 1684–1689.
Hamel N, Tischkowitz M & Foulkes WD (2008) A role for PALB2/FANCN in breast
cancer predisposition? Med Sci (Paris) 24(2): 120–121.
Hanahan D & Weinberg RA (2000) The hallmarks of cancer. Cell 100(1): 57–70.
Hanahan D & Weinberg RA (2011) Hallmarks of cancer: the next generation. Cell 144(5):
646–674.
Hanenberg H, Batish SD, Pollok KE, Vieten L, Verlander PC, Leurs C, Cooper RJ,
Gottsche K, Haneline L, Clapp DW, Lobitz S, Williams DA & Auerbach AD (2002)
Phenotypic correction of primary Fanconi anemia T cells with retroviral vectors as a
diagnostic tool. Exp Hematol 30(5): 410–420.
Harper JW & Elledge SJ (2007) The DNA damage response: ten years after. Mol Cell
28(5): 739–745.
Hayakawa T, Ohtani Y, Hayakawa N, Shinmyozu K, Saito M, Ishikawa F & Nakayama J
(2007) RBP2 is an MRG15 complex component and down-regulates intragenic
histone H3 lysine 4 methylation. Genes Cells 12(6): 811–826.
Hayakawa T, Zhang F, Hayakawa N, Ohtani Y, Shinmyozu K, Nakayama J & Andreassen
PR (2010) MRG15 binds directly to PALB2 and stimulates homology-directed repair
of chromosomal breaks. J Cell Sci 123(Pt 7): 1124–1130.
Hearle N, Schumacher V, Menko FH, Olschwang S, Boardman LA, Gille JJ, Keller JJ,
Westerman AM, Scott RJ, Lim W, Trimbath JD, Giardiello FM, Gruber SB,
Offerhaus GJ, de Rooij FW, Wilson JH, Hansmann A, Moslein G, Royer-Pokora B,
Vogel T, Phillips RK, Spigelman AD & Houlston RS (2006) Frequency and spectrum
of cancers in the Peutz-Jeghers syndrome. Clin Cancer Res 12(10): 3209–3215.
Hebsgaard SM, Korning PG, Tolstrup N, Engelbrecht J, Rouze P & Brunak S (1996)
Splice site prediction in Arabidopsis thaliana pre-mRNA by combining local and
global sequence information. Nucleic Acids Res 24(17): 3439–3452.
Heikkinen K, Karppinen SM, Soini Y, Mäkinen M & Winqvist R (2003) Mutation
screening of Mre11 complex genes: indication of RAD50 involvement in breast and
ovarian cancer susceptibility. J Med Genet 40(12): e131.
Heikkinen K, Mansikka V, Karppinen SM, Rapakko K & Winqvist R (2005) Mutation
analysis of the ATR gene in breast and ovarian cancer families. Breast Cancer Res
7(4): R495–501.
92
Heikkinen K, Rapakko K, Karppinen SM, Erkko H, Knuutila S, Lundan T, Mannermaa A,
Borresen-Dale AL, Borg A, Barkardottir RB, Petrini J & Winqvist R (2006) RAD50
and NBS1 are breast cancer susceptibility genes associated with genomic instability.
Carcinogenesis 27(8): 1593–1599.
Heikkinen T, Kärkkäinen H, Aaltonen K, Milne RL, Heikkilä P, Aittomäki K, Blomqvist C
& Nevanlinna H (2009) The breast cancer susceptibility mutation PALB2 1592delT is
associated with an aggressive tumor phenotype. Clin Cancer Res 15(9): 3214–3222.
Hellebrand H, Sutter C, Honisch E, Gross E, Wappenschmidt B, Schem C, Deissler H,
Ditsch N, Gress V, Kiechle M, Bartram CR, Schmutzler RK, Niederacher D, Arnold
N & Meindl A (2011) Germline mutations in the PALB2 gene are population specific
and occur with low frequencies in familial breast cancer. Hum Mutat 32(6): E2176–88.
Hemminki A, Markie D, Tomlinson I, Avizienyte E, Roth S, Loukola A, Bignell G,
Warren W, Aminoff M, Hoglund P, Jarvinen H, Kristo P, Pelin K, Ridanpaa M,
Salovaara R, Toro T, Bodmer W, Olschwang S, Olsen AS, Stratton MR, de la
Chapelle A & Aaltonen LA (1998) A serine/threonine kinase gene defective in PeutzJeghers syndrome. Nature 391(6663): 184–187.
Hobert JA & Eng C (2009) PTEN hamartoma tumor syndrome: an overview. Genet Med
11(10): 687–694.
Hodgson S,et al. (2006) A Practical Guide to Human Cancer Genetics. Cambridge,
Cambridge University Press.
Hoeijmakers JH (2001) Genome maintenance mechanisms for preventing cancer. Nature
411(6835): 366–374.
Hollestelle A, Wasielewski M, Martens JW & Schutte M (2010) Discovering moderaterisk breast cancer susceptibility genes. Curr Opin Genet Dev 20(3): 268–276.
Hopker K, Hagmann H, Khurshid S, Chen S, Hasskamp P, Seeger-Nukpezah T, Schilberg
K, Heukamp L, Lamkemeyer T, Sos ML, Thomas RK, Lowery D, Roels F, Fischer M,
Liebau MC, Resch U, Kisner T, Rother F, Bartram MP, Muller RU, Fabretti F,
Kurschat P, Schumacher B, Gaestel M, Medema RH, Yaffe MB, Schermer B,
Reinhardt HC & Benzing T (2012) AATF/Che-1 acts as a phosphorylation-dependent
molecular modulator to repress p53-driven apoptosis. EMBO J 31(20): 3961–3975.
Howlett NG, Taniguchi T, Olson S, Cox B, Waisfisz Q, De Die-Smulders C, Persky N,
Grompe M, Joenje H, Pals G, Ikeda H, Fox EA & D'Andrea AD (2002) Biallelic
inactivation of BRCA2 in Fanconi anemia. Science 297(5581): 606–609.
Huang Y & Li L (2013) DNA crosslinking damage and cancer - a tale of friend and foe.
Transl Cancer Res 2(3): 144–154.
Huen MS, Sy SM & Chen J (2010) BRCA1 and its toolbox for the maintenance of genome
integrity. Nat Rev Mol Cell Biol 11(2): 138–148.
Hugh J, Hanson J, Cheang MC, Nielsen TO, Perou CM, Dumontet C, Reed J, Krajewska
M, Treilleux I, Rupin M, Magherini E, Mackey J, Martin M & Vogel C (2009) Breast
cancer subtypes and response to docetaxel in node-positive breast cancer: use of an
immunohistochemical definition in the BCIRG 001 trial. J Clin Oncol 27(8): 1168–
1176.
93
Humar B & Guilford P (2009) Hereditary diffuse gastric cancer: a manifestation of lost
cell polarity. Cancer Sci 100(7): 1151–1157.
Hussain S, Wilson JB, Blom E, Thompson LH, Sung P, Gordon SM, Kupfer GM, Joenje H,
Mathew CG & Jones NJ (2006) Tetratricopeptide-motif-mediated interaction of
FANCG with recombination proteins XRCC3 and BRCA2. DNA Repair (Amst) 5(5):
629–640.
Huusko P, Castren K, Launonen V, Soini Y, Pääkkönen K, Leisti J, Vähäkangas K &
Winqvist R (1999) Germ-line TP53 mutations in Finnish cancer families exhibiting
features of the Li-Fraumeni syndrome and negative for BRCA1 and BRCA2. Cancer
Genet Cytogenet 112(1): 9–14.
Huusko P, Paakkonen K, Launonen V, Pöyhönen M, Blanco G, Kauppila A, Puistola U,
Kiviniemi H, Kujala M, Leisti J & Winqvist R (1998) Evidence of founder mutations
in Finnish BRCA1 and BRCA2 families. Am J Hum Genet 62(6): 1544–1548.
Ishiai M, Kitao H, Smogorzewska A, Tomida J, Kinomura A, Uchida E, Saberi A,
Kinoshita E, Kinoshita-Kikuta E, Koike T, Tashiro S, Elledge SJ & Takata M (2008)
FANCI phosphorylation functions as a molecular switch to turn on the Fanconi
anemia pathway. Nat Struct Mol Biol 15(11): 1138–1146.
Jackson SP & Bartek J (2009) The DNA-damage response in human biology and disease.
Nature 461(7267): 1071–1078.
Jagut M, Mihaila-Bodart L, Molla-Herman A, Alin MF, Lepesant JA & Huynh JR (2013)
A mosaic genetic screen for genes involved in the early steps of Drosophila oogenesis.
G3 (Bethesda) 3(3): 409–425.
Jones S, Hruban RH, Kamiyama M, Borges M, Zhang X, Parsons DW, Lin JC, Palmisano
E, Brune K, Jaffee EM, Iacobuzio-Donahue CA, Maitra A, Parmigiani G, Kern SE,
Velculescu VE, Kinzler KW, Vogelstein B, Eshleman JR, Goggins M & Klein AP
(2009) Exomic sequencing identifies PALB2 as a pancreatic cancer susceptibility
gene. Science 324(5924): 217.
Kaiser J (2012) Human genetics. Genetic influences on disease remain hidden. Science
338(6110): 1016–1017.
Karppinen SM, Heikkinen K, Rapakko K & Winqvist R (2004) Mutation screening of the
BARD1 gene: evidence for involvement of the Cys557Ser allele in hereditary
susceptibility to breast cancer. J Med Genet 41(9): e114.
Karppinen SM, Vuosku J, Heikkinen K, Allinen M & Winqvist R (2003) No evidence of
involvement of germline BACH1 mutations in Finnish breast and ovarian cancer
families. Eur J Cancer 39(3): 366–371.
Kast K, Krause M, Schuler M, Friedrich K, Thamm B, Bier A, Distler W & Kruger S
(2012) Late onset Li-Fraumeni Syndrome with bilateral breast cancer and other
malignancies: case report and review of the literature. BMC Cancer 12: 217.
Kaul D & Mehrotra A (2007) Functional characterization of AATF transcriptome in
human leukemic cells. Mol Cell Biochem 297(1–2): 215–220.
94
Kaurah P, MacMillan A, Boyd N, Senz J, De Luca A, Chun N, Suriano G, Zaor S, Van
Manen L, Gilpin C, Nikkel S, Connolly-Wilson M, Weissman S, Rubinstein WS,
Sebold C, Greenstein R, Stroop J, Yim D, Panzini B, McKinnon W, Greenblatt M,
Wirtzfeld D, Fontaine D, Coit D, Yoon S, Chung D, Lauwers G, Pizzuti A, Vaccaro C,
Redal MA, Oliveira C, Tischkowitz M, Olschwang S, Gallinger S, Lynch H, Green J,
Ford J, Pharoah P, Fernandez B & Huntsman D (2007) Founder and recurrent CDH1
mutations in families with hereditary diffuse gastric cancer. JAMA 297(21): 2360–
2372.
Kee Y & D'Andrea AD (2012) Molecular pathogenesis and clinical management of
Fanconi anemia. J Clin Invest 122(11): 3799–3806.
Kim H & D'Andrea AD (2012) Regulation of DNA cross-link repair by the Fanconi
anemia/BRCA pathway. Genes Dev 26(13): 1393–1408.
Kim JM, Parmar K, Huang M, Weinstock DM, Ruit CA, Kutok JL & D'Andrea AD (2009)
Inactivation of murine Usp1 results in genomic instability and a Fanconi anemia
phenotype. Dev Cell 16(2): 314–320.
Kim Y, Lach FP, Desetty R, Hanenberg H, Auerbach AD & Smogorzewska A (2011)
Mutations of the SLX4 gene in Fanconi anemia. Nat Genet 43(2): 142–146.
Kinzler KW & Vogelstein B (1997) Cancer-susceptibility genes. Gatekeepers and
caretakers. Nature 386(6627): 761, 763.
Kinzler KW & Vogelstein B (1998) Landscaping the cancer terrain. Science 280(5366):
1036–1037.
Kitao H & Takata M (2011) Fanconi anemia: a disorder defective in the DNA damage
response. Int J Hematol 93(4): 417–424.
Knipscheer P, Raschle M, Smogorzewska A, Enoiu M, Ho TV, Scharer OD, Elledge SJ &
Walter JC (2009) The Fanconi anemia pathway promotes replication-dependent DNA
interstrand cross-link repair. Science 326(5960): 1698–1701.
Knudson AG,Jr (1971) Mutation and cancer: statistical study of retinoblastoma. Proc Natl
Acad Sci USA 68(4): 820–823.
Koonin EV, Altschul SF & Bork P (1996) BRCA1 protein products ... Functional motifs..
Nat Genet 13(3): 266–268.
Körkkö J, Annunen S, Pihlajamaa T, Prockop DJ & Ala-Kokko L (1998) Conformation
sensitive gel electrophoresis for simple and accurate detection of mutations:
comparison with denaturing gradient gel electrophoresis and nucleotide sequencing.
Proc Natl Acad Sci USA 95(4): 1681–1685.
Kottemann MC & Smogorzewska A (2013) Fanconi anaemia and the repair of Watson and
Crick DNA crosslinks. Nature 493(7432): 356–363.
Kruyt FA, Abou-Zahr F, Mok H & Youssoufian H (1999) Resistance to mitomycin C
requires direct interaction between the Fanconi anemia proteins FANCA and FANCG
in the nucleus through an arginine-rich domain. J Biol Chem 274(48): 34212–34218.
Kurz EU & Lees-Miller SP (2004) DNA damage-induced activation of ATM and ATMdependent signaling pathways. DNA Repair (Amst) 3(8–9): 889–900.
95
Lakhani SR, Reis-Filho JS, Fulford L, Penault-Llorca F, van der Vijver M, Parry S, Bishop
T, Benitez J, Rivas C, Bignon YJ, Chang-Claude J, Hamann U, Cornelisse CJ,
Devilee P, Beckmann MW, Nestle-Kramling C, Daly PA, Haites N, Varley J, Lalloo F,
Evans G, Maugard C, Meijers-Heijboer H, Klijn JG, Olah E, Gusterson BA, Pilotti S,
Radice P, Scherneck S, Sobol H, Jacquemier J, Wagner T, Peto J, Stratton MR,
McGuffog L, Easton DF & Breast Cancer Linkage Consortium (2005) Prediction of
BRCA1 status in patients with breast cancer using estrogen receptor and basal
phenotype. Clin Cancer Res 11(14): 5175–5180.
Lalloo F & Evans DG (2012) Familial breast cancer. Clin Genet 82(2): 105–114.
Lalloo F, Varley J, Ellis D, Moran A, O'Dair L, Pharoah P, Evans DG & Early Onset
Breast Cancer Study Group (2003) Prediction of pathogenic mutations in patients with
early-onset breast cancer by family history. Lancet 361(9363): 1101–1102.
Lamarche BJ, Orazio NI & Weitzman MD (2010) The MRN complex in double-strand
break repair and telomere maintenance. FEBS Lett 584(17): 3682–3695.
Latif A, Hadfield KD, Roberts SA, Shenton A, Lalloo F, Black GC, Howell A, Evans DG
& Newman WG (2010) Breast cancer susceptibility variants alter risks in familial
disease. J Med Genet 47(2): 126–131.
Lavin MF (2007) ATM and the Mre11 complex combine to recognize and signal DNA
double-strand breaks. Oncogene 26(56): 7749–7758.
Le Calvez-Kelm F, Oliver J, Damiola F, Forey N, Robinot N, Durand G, Voegele C,
Vallee MP, Byrnes G, Registry BC, Hopper JL, Southey MC, Andrulis IL, John EM,
Tavtigian SV & Lesueur F (2012) RAD51 and breast cancer susceptibility: no
evidence for rare variant association in the Breast Cancer Family Registry study.
PLoS One 7(12): e52374.
Leung JW, Wang Y, Fong KW, Huen MS, Li L & Chen J (2012) Fanconi anemia (FA)
binding protein FAAP20 stabilizes FA complementation group A (FANCA) and
participates in interstrand cross-link repair. Proc Natl Acad Sci USA 109(12): 4491–
4496.
Levitus M, Waisfisz Q, Godthelp BC, de Vries Y, Hussain S, Wiegant WW, ElghalbzouriMaghrani E, Steltenpool J, Rooimans MA, Pals G, Arwert F, Mathew CG, Zdzienicka
MZ, Hiom K, De Winter JP & Joenje H (2005) The DNA helicase BRIP1 is defective
in Fanconi anemia complementation group J. Nat Genet 37(9): 934–935.
Levran O, Attwooll C, Henry RT, Milton KL, Neveling K, Rio P, Batish SD, Kalb R,
Velleuer E, Barral S, Ott J, Petrini J, Schindler D, Hanenberg H & Auerbach AD
(2005) The BRCA1-interacting helicase BRIP1 is deficient in Fanconi anemia. Nat
Genet 37(9): 931–933.
Levy-Lahad E (2010) Fanconi anemia and breast cancer susceptibility meet again. Nat
Genet 42(5): 368–369.
Lewis BP, Burge CB & Bartel DP (2005) Conserved seed pairing, often flanked by
adenosines, indicates that thousands of human genes are microRNA targets. Cell
120(1): 15–20.
Li CI, Anderson BO, Daling JR & Moe RE (2003) Trends in incidence rates of invasive
lobular and ductal breast carcinoma. JAMA 289(11): 1421–1424.
96
Li CI, Uribe DJ & Daling JR (2005) Clinical characteristics of different histologic types of
breast cancer. Br J Cancer 93(9): 1046–1052.
Li ML & Greenberg RA (2012) Links between genome integrity and BRCA1 tumor
suppression. Trends Biochem Sci 37(10): 418–424.
Lichtenstein P, Holm NV, Verkasalo PK, Iliadou A, Kaprio J, Koskenvuo M, Pukkala E,
Skytthe A & Hemminki K (2000) Environmental and heritable factors in the causation
of cancer--analyses of cohorts of twins from Sweden, Denmark, and Finland. N Engl J
Med 343(2): 78–85.
Lindahl T & Barnes DE (2000) Repair of endogenous DNA damage. Cold Spring Harb
Symp Quant Biol 65: 127–133.
Lindfors K, Halttunen T, Huotari P, Nupponen N, Vihinen M, Visakorpi T, Mäki M &
Kainulainen H (2000) Identification of novel transcription factor-like gene from
human intestinal cells. Biochem Biophys Res Commun 276(2): 660–666.
Litman R, Peng M, Jin Z, Zhang F, Zhang J, Powell S, Andreassen PR & Cantor SB (2005)
BACH1 is critical for homologous recombination and appears to be the Fanconi
anemia gene product FANCJ. Cancer Cell 8(3): 255–265.
Liu R, Hu T, Li JH, Liang C, Gu WY, Shi XD & Wang HX (2013) Diagnosis of Fanconi
anemia in children with atypical clinical features: a primary study. Chin Med J (Engl)
126(23): 4483–4486.
Livingston DM (2009) Cancer. Complicated supercomplexes. Science 324(5927): 602–603.
Lloyd KM,2nd & Dennis M (1963) Cowden's disease. A possible new symptom complex
with multiple system involvement. Ann Intern Med 58: 136–142.
Lo Ten Foe JR, Rooimans MA, Bosnoyan-Collins L, Alon N, Wijker M, Parker L,
Lightfoot J, Carreau M, Callen DF, Savoia A, Cheng NC, van Berkel CG, Strunk MH,
Gille JJ, Pals G, Kruyt FA, Pronk JC, Arwert F, Buchwald M & Joenje H (1996)
Expression cloning of a cDNA for the major Fanconi anaemia gene, FAA. Nat Genet
14(3): 320–323.
Loeb LA (1991) Mutator phenotype may be required for multistage carcinogenesis. Cancer
Res 51(12): 3075–3079.
Loeb LA (2001) A mutator phenotype in cancer. Cancer Res 61(8): 3230–3239.
Lord CJ & Ashworth A (2012) The DNA damage response and cancer therapy. Nature
481(7381): 287–294.
Loveday C, Turnbull C, Ruark E, Xicola RM, Ramsay E, Hughes D, Warren-Perry M,
Snape K, Breast Cancer Susceptibility Collaboration (UK), Eccles D, Evans DG, Gore
M, Renwick A, Seal S, Antoniou AC & Rahman N (2012) Germline RAD51C
mutations confer susceptibility to ovarian cancer. Nat Genet 44(5): 475–6; author
reply 476.
Lukas J, Lukas C & Bartek J (2004) Mammalian cell cycle checkpoints: signalling
pathways and their organization in space and time. DNA Repair (Amst) 3(8–9): 997–
1007.
Malkin D, Li FP, Strong LC, Fraumeni JF,Jr, Nelson CE, Kim DH, Kassel J, Gryka MA,
Bischoff FZ & Tainsky MA (1990) Germ line p53 mutations in a familial syndrome
of breast cancer, sarcomas, and other neoplasms. Science 250(4985): 1233–1238.
97
Mallory SB (1995) Cowden syndrome (multiple hamartoma syndrome). Dermatol Clin
13(1): 27–31.
Manolio TA, Collins FS, Cox NJ, Goldstein DB, Hindorff LA, Hunter DJ, McCarthy MI,
Ramos EM, Cardon LR, Chakravarti A, Cho JH, Guttmacher AE, Kong A, Kruglyak
L, Mardis E, Rotimi CN, Slatkin M, Valle D, Whittemore AS, Boehnke M, Clark AG,
Eichler EE, Gibson G, Haines JL, Mackay TF, McCarroll SA & Visscher PM (2009)
Finding the missing heritability of complex diseases. Nature 461(7265): 747–753.
Marsh D & Zori R (2002) Genetic insights into familial cancers-- update and recent
discoveries. Cancer Lett 181(2): 125–164.
Marshall CJ (1991) Tumor suppressor genes. Cell 64(2): 313–326.
Martrat G, Maxwell CM, Tominaga E, Porta-de-la-Riva M, Bonifaci N, Gomez-Baldo L,
Bogliolo M, Lazaro C, Blanco I, Brunet J, Aguilar H, Fernandez-Rodriguez J, Seal S,
Renwick A, Rahman N, Kuhl J, Neveling K, Schindler D, Ramirez MJ, Castella M,
Hernandez G, EMBRACE, Easton DF, Peock S, Cook M, Oliver CT, Frost D, Platte
R, Evans DG, Lalloo F, Eeles R, Izatt L, Chu C, Davidson R, Ong KR, Cook J,
Douglas F, Hodgson S, Brewer C, Morrison PJ, Porteous M, Peterlongo P, Manoukian
S, Peissel B, Zaffaroni D, Roversi G, Barile M, Viel A, Pasini B, Ottini L, Putignano
AL, Savarese A, Bernard L, Radice P, Healey S, Spurdle A, Chen X, Beesley J,
kConFab, Rookus MA, Verhoef S, Tilanus-Linthorst MA, Vreeswijk MP, Asperen CJ,
Bodmer D, Ausems MG, van Os TA, Blok MJ, Meijers-Heijboer HE, Hogervorst FB,
HEBON, Goldgar DE, Buys S, John EM, Miron A, Southey M, Daly MB, BCFR,
SWE-BRCA, Harbst K, Borg A, Rantala J, Barbany-Bustinza G, Ehrencrona H,
Stenmark-Askmalm M, Kaufman B, Laitman Y, Milgrom R, Friedman E, Domchek
SM, Nathanson KL, Rebbeck TR, Johannsson OT, Couch FJ, Wang X, Fredericksen
Z, Cuadras D, Moreno V, Pientka FK, Depping R, Caldes T, Osorio A, Benitez J,
Bueren J, Heikkinen T, Nevanlinna H, Hamann U, Torres D, Caligo MA, Godwin AK,
Imyanitov EN, Janavicius R, GEMO Study Collaborators, Sinilnikova OM, StoppaLyonnet D, Mazoyer S, Verny-Pierre C, Castera L, de Pauw A, Bignon YJ,
Uhrhammer N, Peyrat JP, Vennin P, Ferrer SF, Collonge-Rame MA, Mortemousque I,
McGuffog L, Chenevix-Trench G, Pereira-Smith OM, Antoniou AC, Ceron J,
Tominaga K, Surralles J & Pujana MA (2011) Exploring the link between MORF4L1
and risk of breast cancer. Breast Cancer Res 13(2): R40.
Matsuoka S, Huang M & Elledge SJ (1998) Linkage of ATM to cell cycle regulation by
the Chk2 protein kinase. Science 282(5395): 1893–1897.
98
Mavaddat N, Barrowdale D, Andrulis IL, Domchek SM, Eccles D, Nevanlinna H, Ramus
SJ, Spurdle A, Robson M, Sherman M, Mulligan AM, Couch FJ, Engel C, McGuffog
L, Healey S, Sinilnikova OM, Southey MC, Terry MB, Goldgar D, O'Malley F, John
EM, Janavicius R, Tihomirova L, Hansen TV, Nielsen FC, Osorio A, Stavropoulou A,
Benitez J, Manoukian S, Peissel B, Barile M, Volorio S, Pasini B, Dolcetti R,
Putignano AL, Ottini L, Radice P, Hamann U, Rashid MU, Hogervorst FB, Kriege M,
van der Luijt RB, HEBON, Peock S, Frost D, Evans DG, Brewer C, Walker L, Rogers
MT, Side LE, Houghton C, EMBRACE, Weaver J, Godwin AK, Schmutzler RK,
Wappenschmidt B, Meindl A, Kast K, Arnold N, Niederacher D, Sutter C, Deissler H,
Gadzicki D, Preisler-Adams S, Varon-Mateeva R, Schonbuchner I, Gevensleben H,
Stoppa-Lyonnet D, Belotti M, Barjhoux L, GEMO Study Collaborators, Isaacs C,
Peshkin BN, Caldes T, de la Hoya M, Canadas C, Heikkinen T, Heikkila P, Aittomaki
K, Blanco I, Lazaro C, Brunet J, Agnarsson BA, Arason A, Barkardottir RB, Dumont
M, Simard J, Montagna M, Agata S, D'Andrea E, Yan M, Fox S, kConFab
Investigators, Rebbeck TR, Rubinstein W, Tung N, Garber JE, Wang X, Fredericksen
Z, Pankratz VS, Lindor NM, Szabo C, Offit K, Sakr R, Gaudet MM, Singer CF, Tea
MK, Rappaport C, Mai PL, Greene MH, Sokolenko A, Imyanitov E, Toland AE,
Senter L, Sweet K, Thomassen M, Gerdes AM, Kruse T, Caligo M, Aretini P, Rantala
J, von Wachenfeld A, Henriksson K, SWE-BRCA Collaborators, Steele L, Neuhausen
SL, Nussbaum R, Beattie M, Odunsi K, Sucheston L, Gayther SA, Nathanson K,
Gross J, Walsh C, Karlan B, Chenevix-Trench G, Easton DF, Antoniou AC &
Consortium of Investigators of Modifiers of BRCA1/2 (2012) Pathology of breast and
ovarian cancers among BRCA1 and BRCA2 mutation carriers: results from the
Consortium of Investigators of Modifiers of BRCA1/2 (CIMBA). Cancer Epidemiol
Biomarkers Prev 21(1): 134–147.
Mavaddat N, Pharoah PD, Blows F, Driver KE, Provenzano E, Thompson D, Macinnis RJ,
Shah M, SEARCH Team, Easton DF & Antoniou AC (2010) Familial relative risks
for breast cancer by pathological subtype: a population-based cohort study. Breast
Cancer Res 12(1): R10.
Maxwell KN & Nathanson KL (2013) Common breast cancer risk variants in the postCOGS era: a comprehensive review. Breast Cancer Res 15(6): 212.
McClellan J & King MC (2010) Genetic heterogeneity in human disease. Cell 141(2):
210–217.
McKinnon PJ & Caldecott KW (2007) DNA strand break repair and human genetic disease.
Annu Rev Genomics Hum Genet 8: 37–55.
99
Meijers-Heijboer H, van den Ouweland A, Klijn J, Wasielewski M, de Snoo A, Oldenburg
R, Hollestelle A, Houben M, Crepin E, van Veghel-Plandsoen M, Elstrodt F, van
Duijn C, Bartels C, Meijers C, Schutte M, McGuffog L, Thompson D, Easton D,
Sodha N, Seal S, Barfoot R, Mangion J, Chang-Claude J, Eccles D, Eeles R, Evans
DG, Houlston R, Murday V, Narod S, Peretz T, Peto J, Phelan C, Zhang HX, Szabo C,
Devilee P, Goldgar D, Futreal PA, Nathanson KL, Weber B, Rahman N, Stratton MR
& CHEK2-Breast Cancer Consortium (2002) Low-penetrance susceptibility to breast
cancer due to CHEK2(*)1100delC in noncarriers of BRCA1 or BRCA2 mutations.
Nat Genet 31(1): 55–59.
Meindl A (2009) Identification of Novel Susceptibility Genes for Breast Cancer - GenomeWide Association Studies or Evaluation of Candidate Genes? Breast Care (Basel) 4(2):
93–99.
Meindl A, Hellebrand H, Wiek C, Erven V, Wappenschmidt B, Niederacher D, Freund M,
Lichtner P, Hartmann L, Schaal H, Ramser J, Honisch E, Kubisch C, Wichmann HE,
Kast K, Deissler H, Engel C, Muller-Myhsok B, Neveling K, Kiechle M, Mathew CG,
Schindler D, Schmutzler RK & Hanenberg H (2010) Germline mutations in breast and
ovarian cancer pedigrees establish RAD51C as a human cancer susceptibility gene.
Nat Genet 42(5): 410–414.
100
Michailidou K, Hall P, Gonzalez-Neira A, Ghoussaini M, Dennis J, Milne RL, Schmidt
MK, Chang-Claude J, Bojesen SE, Bolla MK, Wang Q, Dicks E, Lee A, Turnbull C,
Rahman N, Breast and Ovarian Cancer Susceptibility Collaboration, Fletcher O, Peto
J, Gibson L, Dos Santos Silva I, Nevanlinna H, Muranen TA, Aittomaki K, Blomqvist
C, Czene K, Irwanto A, Liu J, Waisfisz Q, Meijers-Heijboer H, Adank M, Hereditary
Breast and Ovarian Cancer Research Group Netherlands (HEBON), van der Luijt RB,
Hein R, Dahmen N, Beckman L, Meindl A, Schmutzler RK, Muller-Myhsok B,
Lichtner P, Hopper JL, Southey MC, Makalic E, Schmidt DF, Uitterlinden AG,
Hofman A, Hunter DJ, Chanock SJ, Vincent D, Bacot F, Tessier DC, Canisius S,
Wessels LF, Haiman CA, Shah M, Luben R, Brown J, Luccarini C, Schoof N,
Humphreys K, Li J, Nordestgaard BG, Nielsen SF, Flyger H, Couch FJ, Wang X,
Vachon C, Stevens KN, Lambrechts D, Moisse M, Paridaens R, Christiaens MR,
Rudolph A, Nickels S, Flesch-Janys D, Johnson N, Aitken Z, Aaltonen K, Heikkinen
T, Broeks A, Veer LJ, van der Schoot CE, Guenel P, Truong T, Laurent-Puig P,
Menegaux F, Marme F, Schneeweiss A, Sohn C, Burwinkel B, Zamora MP, Perez JI,
Pita G, Alonso MR, Cox A, Brock IW, Cross SS, Reed MW, Sawyer EJ, Tomlinson I,
Kerin MJ, Miller N, Henderson BE, Schumacher F, Le Marchand L, Andrulis IL,
Knight JA, Glendon G, Mulligan AM, kConFab Investigators, Australian Ovarian
Cancer Study Group, Lindblom A, Margolin S, Hooning MJ, Hollestelle A, van den
Ouweland AM, Jager A, Bui QM, Stone J, Dite GS, Apicella C, Tsimiklis H, Giles
GG, Severi G, Baglietto L, Fasching PA, Haeberle L, Ekici AB, Beckmann MW,
Brenner H, Muller H, Arndt V, Stegmaier C, Swerdlow A, Ashworth A, Orr N, Jones
M, Figueroa J, Lissowska J, Brinton L, Goldberg MS, Labreche F, Dumont M,
Winqvist R, Pylkas K, Jukkola-Vuorinen A, Grip M, Brauch H, Hamann U, Bruning
T, GENICA (Gene Environment Interaction and Breast Cancer in Germany) Network,
Radice P, Peterlongo P, Manoukian S, Bonanni B, Devilee P, Tollenaar RA, Seynaeve
C, van Asperen CJ, Jakubowska A, Lubinski J, Jaworska K, Durda K, Mannermaa A,
Kataja V, Kosma VM, Hartikainen JM, Bogdanova NV, Antonenkova NN, Dork T,
Kristensen VN, Anton-Culver H, Slager S, Toland AE, Edge S, Fostira F, Kang D,
Yoo KY, Noh DY, Matsuo K, Ito H, Iwata H, Sueta A, Wu AH, Tseng CC, Van Den
Berg D, Stram DO, Shu XO, Lu W, Gao YT, Cai H, Teo SH, Yip CH, Phuah SY,
Cornes BK, Hartman M, Miao H, Lim WY, Sng JH, Muir K, Lophatananon A,
Stewart-Brown S, Siriwanarangsan P, Shen CY, Hsiung CN, Wu PE, Ding SL,
Sangrajrang S, Gaborieau V, Brennan P, McKay J, Blot WJ, Signorello LB, Cai Q,
Zheng W, Deming-Halverson S, Shrubsole M, Long J, Simard J, Garcia-Closas M,
Pharoah PD, Chenevix-Trench G, Dunning AM, Benitez J & Easton DF (2013) Largescale genotyping identifies 41 new loci associated with breast cancer risk. Nat Genet
45(4): 353–61, 361e1–2.
Miki Y, Swensen J, Shattuck-Eidens D, Futreal PA, Harshman K, Tavtigian S, Liu Q,
Cochran C, Bennett LM & Ding W (1994) A strong candidate for the breast and
ovarian cancer susceptibility gene BRCA1. Science 266(5182): 66–71.
101
Milne RL, John EM, Knight JA, Dite GS, Southey MC, Giles GG, Apicella C, West DW,
Andrulis IL, Whittemore AS & Hopper JL (2011) The potential value of sibling
controls compared with population controls for association studies of lifestyle-related
risk factors: an example from the Breast Cancer Family Registry. Int J Epidemiol
40(5): 1342–1354.
Mirchandani KD & D'Andrea AD (2006) The Fanconi anemia/BRCA pathway: a
coordinator of cross-link repair. Exp Cell Res 312(14): 2647–2653.
Mirzayans R, Andrais B, Scott A & Murray D (2012) New insights into p53 signaling and
cancer cell response to DNA damage: implications for cancer therapy. J Biomed
Biotechnol 2012: 170325.
Mouridsen HT (1993) Adjuvant systemic therapy in breast cancer. Eur J Cancer 29A(4):
595–598.
Muller PA & Vousden KH (2013) P53 Mutations in Cancer. Nat Cell Biol 15(1): 2–8.
Nakayama J, Xiao G, Noma K, Malikzay A, Bjerling P, Ekwall K, Kobayashi R & Grewal
SI (2003) Alp13, an MRG family protein, is a component of fission yeast Clr6 histone
deacetylase required for genomic integrity. EMBO J 22(11): 2776–2787.
Narod SA & Foulkes WD (2004) BRCA1 and BRCA2: 1994 and beyond. Nat Rev Cancer
4(9): 665–676.
Nathanson KL & Weber BL (2001) "Other" breast cancer susceptibility genes: searching
for more holy grail. Hum Mol Genet 10(7): 715–720.
National Cancer Institute: SEER Surveillance epidemiology and end results: Cancer stat
fact sheets: breast. URI: http://seer.cancer.gov/csr/1975_2010/results_single/
sect_01_table.12_2pgs.pdf.
National Institute for Health and Welfare (THL) Terveyden ja hyvinvoinnin laitos. Tietoa
rintasyövän seulonnasta. URI: http://www.thl.fi/fi_FI/web/fi/aiheet/tietopaketit/
seulonnat/rintasyovan_seulonta.
Nelen MR, Padberg GW, Peeters EA, Lin AY, van den Helm B, Frants RR, Coulon V,
Goldstein AM, van Reen MM, Easton DF, Eeles RA, Hodgsen S, Mulvihill JJ,
Murday VA, Tucker MA, Mariman EC, Starink TM, Ponder BA, Ropers HH, Kremer
H, Longy M & Eng C (1996) Localization of the gene for Cowden disease to
chromosome 10q22–23. Nat Genet 13(1): 114–116.
Niedernhofer LJ, Lalai AS & Hoeijmakers JH (2005) Fanconi anemia (cross)linked to
DNA repair. Cell 123(7): 1191–1198.
Niedzwiedz W, Mosedale G, Johnson M, Ong CY, Pace P & Patel KJ (2004) The Fanconi
anaemia gene FANCC promotes homologous recombination and error-prone DNA
repair. Mol Cell 15(4): 607–620.
Nikkilä J, Coleman KA, Morrissey D, Pylkäs K, Erkko H, Messick TE, Karppinen SM,
Amelina A, Winqvist R & Greenberg RA (2009) Familial breast cancer screening
reveals an alteration in the RAP80 UIM domain that impairs DNA damage response
function. Oncogene 28(16): 1843–1852.
Nikkilä J, Parplys AC, Pylkäs K, Bose M, Huo Y, Borgmann K, Rapakko K, Nieminen P,
Xia B, Pospiech H & Winqvist R (2013) Heterozygous mutations in PALB2 cause
DNA replication and damage response defects. Nat Commun 4: 2578.
102
Novak DJ, Sabbaghian N, Maillet P, Chappuis PO, Foulkes WD & Tischkowitz M (2009)
Analysis of the genes coding for the BRCA1-interacting proteins, RAP80 and
Abraxas (CCDC98), in high-risk, non-BRCA1/2, multiethnic breast cancer cases.
Breast Cancer Res Treat 117(2): 453–459.
Nowell PC (1976) The clonal evolution of tumor cell populations. Science 194(4260): 23–
28.
Oliver AW, Swift S, Lord CJ, Ashworth A & Pearl LH (2009) Structural basis for
recruitment of BRCA2 by PALB2. EMBO Rep 10(9): 990–996.
Oricchio E, Saladino C, Iacovelli S, Soddu S & Cundari E (2006) ATM is activated by
default in mitosis, localizes at centrosomes and monitors mitotic spindle integrity. Cell
Cycle 5(1): 88–92.
Osorio A, Barroso A, Garcia MJ, Martinez-Delgado B, Urioste M & Benitez J (2009)
Evaluation of the BRCA1 interacting genes RAP80 and CCDC98 in familial breast
cancer susceptibility. Breast Cancer Res Treat 113(2): 371–376.
Osorio A, Endt D, Fernandez F, Eirich K, de la Hoya M, Schmutzler R, Caldes T, Meindl
A, Schindler D & Benitez J (2012) Predominance of pathogenic missense variants in
the RAD51C gene occurring in breast and ovarian cancer families. Hum Mol Genet
21(13): 2889–2898.
Page G, Lodige I, Kogel D & Scheidtmann KH (1999) AATF, a novel transcription factor
that interacts with Dlk/ZIP kinase and interferes with apoptosis. FEBS Lett 462(1–2):
187–191.
Palmeri S, Rufa A, Pucci B, Santarnecchi E, Malandrini A, Stromillo ML, Mandala M,
Rosini F, De Stefano N & Federico A (2013) Clinical course of two Italian siblings
with ataxia-telangiectasia-like disorder. Cerebellum 12(4): 596–599.
Parkin DM (2004) International variation. Oncogene 23(38): 6329–6340.
Pasche B & Yi N (2010) Candidate gene association studies: successes and failures. Curr
Opin Genet Dev 20(3): 257–261.
Passananti C, Floridi A & Fanciulli M (2007) Che-1/AATF, a multivalent adaptor
connecting transcriptional regulation, checkpoint control, and apoptosis. Biochem Cell
Biol 85(4): 477–483.
Patel KJ & Joenje H (2007) Fanconi anemia and DNA replication repair. DNA Repair
(Amst) 6(7): 885–890.
Patnala R, Clements J & Batra J (2013) Candidate gene association studies: a
comprehensive guide to useful in silico tools. BMC Genet 14: 39–2156–14–39.
Payne SR & Kemp CJ (2005) Tumor suppressor genetics. Carcinogenesis 26(12): 2031–
2045.
Pellegrini L, Yu DS, Lo T, Anand S, Lee M, Blundell TL & Venkitaraman AR (2002)
Insights into DNA recombination from the structure of a RAD51-BRCA2 complex.
Nature 420(6913): 287–293.
Pelttari LM, Heikkinen T, Thompson D, Kallioniemi A, Schleutker J, Holli K, Blomqvist
C, Aittomäki K, Butzow R & Nevanlinna H (2011) RAD51C is a susceptibility gene
for ovarian cancer. Hum Mol Genet 20(16): 3278–3288.
103
Pelttari LM, Nurminen R, Gylfe A, Aaltonen LA, Schleutker J & Nevanlinna H (2012)
Screening of Finnish RAD51C founder mutations in prostate and colorectal cancer
patients. BMC Cancer 12: 552.
Pena AN & Pereira-Smith OM (2007) The role of the MORF/MRG family of genes in cell
growth, differentiation, DNA repair, and thereby aging. Ann N Y Acad Sci 1100:
299–305.
Peng M, Litman R, Jin Z, Fong G & Cantor SB (2006) BACH1 is a DNA repair protein
supporting BRCA1 damage response. Oncogene 25(15): 2245–2253.
Pennington KP & Swisher EM (2012) Hereditary ovarian cancer: beyond the usual
suspects. Gynecol Oncol 124(2): 347–353.
Penrose LS, Mackenzie HJ & Karn MN (1948) A genetical study of human mammary
cancer. Ann Eugen 14 Pt(3): 234–266.
Peto J (2001) Cancer epidemiology in the last century and the next decade. Nature
411(6835): 390–395.
Peto J, Collins N, Barfoot R, Seal S, Warren W, Rahman N, Easton DF, Evans C, Deacon J
& Stratton MR (1999) Prevalence of BRCA1 and BRCA2 gene mutations in patients
with early-onset breast cancer. J Natl Cancer Inst 91(11): 943–949.
Petrakis NL (1977) Genetic factors in the etiology of breast cancer. Cancer 39(6 Suppl):
2709–2715.
Pharoah PD, Antoniou A, Bobrow M, Zimmern RL, Easton DF & Ponder BA (2002)
Polygenic susceptibility to breast cancer and implications for prevention. Nat Genet
31(1): 33–36.
Pharoah PD, Antoniou AC, Easton DF & Ponder BA (2008) Polygenes, risk prediction,
and targeted prevention of breast cancer. N Engl J Med 358(26): 2796–2803.
Pharoah PD, Guilford P, Caldas C & International Gastric Cancer Linkage Consortium
(2001) Incidence of gastric cancer and breast cancer in CDH1 (E-cadherin) mutation
carriers from hereditary diffuse gastric cancer families. Gastroenterology 121(6):
1348–1353.
Polo SE & Jackson SP (2011) Dynamics of DNA damage response proteins at DNA
breaks: a focus on protein modifications. Genes Dev 25(5): 409–433.
Pomerantz MM, Ahmadiyeh N, Jia L, Herman P, Verzi MP, Doddapaneni H, Beckwith
CA, Chan JA, Hills A, Davis M, Yao K, Kehoe SM, Lenz HJ, Haiman CA, Yan C,
Henderson BE, Frenkel B, Barretina J, Bass A, Tabernero J, Baselga J, Regan MM,
Manak JR, Shivdasani R, Coetzee GA & Freedman ML (2009) The 8q24 cancer risk
variant rs6983267 shows long-range interaction with MYC in colorectal cancer. Nat
Genet 41(8): 882–884.
Ponder BA (2001) Cancer genetics. Nature 411(6835): 336–341.
Poumpouridou N & Kroupis C (2011) Hereditary breast cancer: beyond BRCA genetic
analysis; PALB2 emerges. Clin Chem Lab Med 50(3): 423–434.
Prasher JM, Lalai AS, Heijmans-Antonissen C, Ploemacher RE, Hoeijmakers JH, Touw IP
& Niedernhofer LJ (2005) Reduced hematopoietic reserves in DNA interstrand
crosslink repair-deficient Ercc1-/- mice. EMBO J 24(4): 861–871.
104
Price BD & D'Andrea AD (2013) Chromatin remodeling at DNA double-strand breaks.
Cell 152(6): 1344–1354.
Pylkäs K, Tommiska J, Syrjäkoski K, Kere J, Gatei M, Waddell N, Allinen M, Karppinen
SM, Rapakko K, Kääriäinen H, Aittomäki K, Blomqvist C, Mustonen A, Holli K,
Khanna KK, Kallioniemi OP, Nevanlinna H & Winqvist R (2007) Evaluation of the
role of Finnish ataxia-telangiectasia mutations in hereditary predisposition to breast
cancer. Carcinogenesis 28(5): 1040–1045.
Rafnar T, Gudbjartsson DF, Sulem P, Jonasdottir A, Sigurdsson A, Jonasdottir A,
Besenbacher S, Lundin P, Stacey SN, Gudmundsson J, Magnusson OT, le Roux L,
Orlygsdottir G, Helgadottir HT, Johannsdottir H, Gylfason A, Tryggvadottir L,
Jonasson JG, de Juan A, Ortega E, Ramon-Cajal JM, Garcia-Prats MD, Mayordomo C,
Panadero A, Rivera F, Aben KK, van Altena AM, Massuger LF, Aavikko M, Kujala
PM, Staff S, Aaltonen LA, Olafsdottir K, Björnsson J, Kong A, Salvarsdottir A,
Saemundsson H, Olafsson K, Benediktsdottir KR, Gulcher J, Masson G, Kiemeney
LA, Mayordomo JI, Thorsteinsdottir U & Stefansson K (2011) Mutations in BRIP1
confer high risk of ovarian cancer. Nat Genet 43(11): 1104–1107.
Rahman N, Seal S, Thompson D, Kelly P, Renwick A, Elliott A, Reid S, Spanova K,
Barfoot R, Chagtai T, Jayatilake H, McGuffog L, Hanks S, Evans DG, Eccles D,
Breast Cancer Susceptibility Collaboration (UK), Easton DF & Stratton MR (2007)
PALB2, which encodes a BRCA2-interacting protein, is a breast cancer susceptibility
gene. Nat Genet 39(2): 165–167.
Ralhan R, Kaur J, Kreienberg R & Wiesmuller L (2007) Links between DNA double
strand break repair and breast cancer: accumulating evidence from both familial and
nonfamilial cases. Cancer Lett 248(1): 1–17.
Ramensky V, Bork P & Sunyaev S (2002) Human non-synonymous SNPs: server and
survey. Nucleic Acids Res 30(17): 3894–3900.
Rapakko K, Allinen M, Syrjäkoski K, Vahteristo P, Huusko P, Vähäkangas K, Eerola H,
Kainu T, Kallioniemi OP, Nevanlinna H & Winqvist R (2001) Germline TP53
alterations in Finnish breast cancer families are rare and occur at conserved mutationprone sites. Br J Cancer 84(1): 116–119.
Rasmussen SA & Friedman JM (2000) NF1 gene and neurofibromatosis 1. Am J
Epidemiol 151(1): 33–40.
Read
A
&
Donnai
D
(2001)
New
Clinical
Genetics
2nd Edition. Banbury, UK, Scion Publishing.
Reed GH & Wittwer CT (2004) Sensitivity and specificity of single-nucleotide
polymorphism scanning by high-resolution melting analysis. Clin Chem 50(10):
1748–1754.
Reese MG, Eeckman FH, Kulp D & Haussler D (1997) Improved splice site detection in
Genie. J Comput Biol 4(3): 311–323.
105
Reid S, Renwick A, Seal S, Baskcomb L, Barfoot R, Jayatilake H, Pritchard-Jones K,
Stratton MR, Ridolfi-Luthy A, Rahman N, Breast Cancer Susceptibility Collaboration
(UK) & Familial Wilms Tumour Collaboration (2005) Biallelic BRCA2 mutations are
associated with multiple malignancies in childhood including familial Wilms tumour.
J Med Genet 42(2): 147–151.
Reid S, Schindler D, Hanenberg H, Barker K, Hanks S, Kalb R, Neveling K, Kelly P, Seal
S, Freund M, Wurm M, Batish SD, Lach FP, Yetgin S, Neitzel H, Ariffin H,
Tischkowitz M, Mathew CG, Auerbach AD & Rahman N (2007) Biallelic mutations
in PALB2 cause Fanconi anemia subtype FA-N and predispose to childhood cancer.
Nat Genet 39(2): 162–164.
Renwick A, Thompson D, Seal S, Kelly P, Chagtai T, Ahmed M, North B, Jayatilake H,
Barfoot R, Spanova K, McGuffog L, Evans DG, Eccles D, Breast Cancer
Susceptibility Collaboration (UK), Easton DF, Stratton MR & Rahman N (2006)
ATM mutations that cause ataxia-telangiectasia are breast cancer susceptibility alleles.
Nat Genet 38(8): 873–875.
Ripperger T, Gadzicki D, Meindl A & Schlegelberger B (2009) Breast cancer
susceptibility: current knowledge and implications for genetic counselling. Eur J Hum
Genet 17(6): 722–731.
Romero A, Perez-Segura P, Tosar A, Garcia-Saenz JA, Diaz-Rubio E, Caldes T & de la
Hoya M (2011) A HRM-based screening method detects RAD51C germ-line
deleterious mutations in Spanish breast and ovarian cancer families. Breast Cancer
Res Treat 129(3): 939–946.
Romick-Rosendale LE, Lui VW, Grandis JR & Wells SI (2013) The Fanconi anemia
pathway: repairing the link between DNA damage and squamous cell carcinoma.
Mutat Res 743–744: 78–88.
Rosen EM (2013) BRCA1 in the DNA damage response and at telomeres. Front Genet 4:
85.
Rosenberg PS, Tamary H & Alter BP (2011) How high are carrier frequencies of rare
recessive syndromes? Contemporary estimates for Fanconi Anemia in the United
States and Israel. Am J Med Genet A 155A(8): 1877–1883.
Ross JS & Hortobagyi GN (2005) Molecular Oncology of breast caner. Mississaunga,
Canada, Jones and Bartlett Publishers.
Rouzier R, Perou CM, Symmans WF, Ibrahim N, Cristofanilli M, Anderson K, Hess KR,
Stec J, Ayers M, Wagner P, Morandi P, Fan C, Rabiul I, Ross JS, Hortobagyi GN &
Pusztai L (2005) Breast cancer molecular subtypes respond differently to preoperative
chemotherapy. Clin Cancer Res 11(16): 5678–5685.
Roy R, Chun J & Powell SN (2011) BRCA1 and BRCA2: different roles in a common
pathway of genome protection. Nat Rev Cancer 12(1): 68–78.
Saar K, Schindler D, Wegner RD, Reis A, Wienker TF, Hoehn H, Joenje H, Sperling K &
Digweed M (1998) Localisation of a Fanconi anaemia gene to chromosome 9p. Eur J
Hum Genet 6(5): 501–508.
106
Savino M, Ianzano L, Strippoli P, Ramenghi U, Arslanian A, Bagnara GP, Joenje H,
Zelante L & Savoia A (1997) Mutations of the Fanconi anemia group A gene (FAA)
in Italian patients. Am J Hum Genet 61(6): 1246–1253.
Savitsky K, Bar-Shira A, Gilad S, Rotman G, Ziv Y, Vanagaite L, Tagle DA, Smith S,
Uziel T, Sfez S, Ashkenazi M, Pecker I, Frydman M, Harnik R, Patanjali SR,
Simmons A, Clines GA, Sartiel A, Gatti RA, Chessa L, Sanal O, Lavin MF, Jaspers
NG, Taylor AM, Arlett CF, Miki T, Weissman SM, Lovett M, Collins FS & Shiloh Y
(1995) A single ataxia telangiectasia gene with a product similar to PI-3 kinase.
Science 268(5218): 1749–1753.
Schlacher K, Wu H & Jasin M (2012) A distinct replication fork protection pathway
connects Fanconi anemia tumor suppressors to RAD51-BRCA1/2. Cancer Cell 22(1):
106–116.
Schouten JP, McElgunn CJ, Waaijer R, Zwijnenburg D, Diepvens F & Pals G (2002)
Relative quantification of 40 nucleic acid sequences by multiplex ligation-dependent
probe amplification. Nucleic Acids Res 30(12): e57.
Scully R (2000) Role of BRCA gene dysfunction in breast and ovarian cancer
predisposition. Breast Cancer Res 2(5): 324–330.
Seal S, Barfoot R, Jayatilake H, Smith P, Renwick A, Bascombe L, McGuffog L, Evans
DG, Eccles D, Easton DF, Stratton MR, Rahman N & Breast Cancer Susceptibility
Collaboration (2003) Evaluation of Fanconi Anemia genes in familial breast cancer
predisposition. Cancer Res 63(24): 8596–8599.
Seal S, Thompson D, Renwick A, Elliott A, Kelly P, Barfoot R, Chagtai T, Jayatilake H,
Ahmed M, Spanova K, North B, McGuffog L, Evans DG, Eccles D, Breast Cancer
Susceptibility Collaboration (UK), Easton DF, Stratton MR & Rahman N (2006)
Truncating mutations in the Fanconi anemia J gene BRIP1 are low-penetrance breast
cancer susceptibility alleles. Nat Genet 38(11): 1239–1241.
Seemanova E (1990) An increased risk for malignant neoplasms in heterozygotes for a
syndrome of microcephaly, normal intelligence, growth retardation, remarkable facies,
immunodeficiency and chromosomal instability. Mutat Res 238(3): 321–324.
Seemanova E, Jarolim P, Seeman P, Varon R, Digweed M, Swift M & Sperling K (2007)
Cancer risk of heterozygotes with the NBN founder mutation. J Natl Cancer Inst
99(24): 1875–1880.
Segev Y, Pal T, Rosen B, McLaughlin JR, Sellers TA, Risch HA, Zhang S, Ping S, Narod
SA & Schildkraut J (2013) Risk factors for ovarian cancers with and without
microsatellite instability. Int J Gynecol Cancer 23(6): 1010–1015.
Sengupta S & Harris CC (2005) p53: traffic cop at the crossroads of DNA repair and
recombination. Nat Rev Mol Cell Biol 6(1): 44–55.
Seshadri R, Kutlaca RJ, Trainor K, Matthews C & Morley AA (1987) Mutation rate of
normal and malignant human lymphocytes. Cancer Res 47(2): 407–409.
Shanley S, Fung C, Milliken J, Leary J, Barnetson R, Schnitzler M & Kirk J (2009) Breast
cancer immunohistochemistry can be useful in triage of some HNPCC families. Fam
Cancer 8(3): 251–255.
107
Sharan SK, Morimatsu M, Albrecht U, Lim DS, Regel E, Dinh C, Sands A, Eichele G,
Hasty P & Bradley A (1997) Embryonic lethality and radiation hypersensitivity
mediated by Rad51 in mice lacking Brca2. Nature 386(6627): 804–810.
Sharif S, Moran A, Huson SM, Iddenden R, Shenton A, Howard E & Evans DG (2007)
Women with neurofibromatosis 1 are at a moderately increased risk of developing
breast cancer and should be considered for early screening. J Med Genet 44(8): 481–
484.
Sharma M (2013) Apoptosis-antagonizing transcription factor (AATF) gene silencing: role
in induction of apoptosis and down-regulation of estrogen receptor in breast cancer
cells. Biotechnol Lett 35(10): 1561–1570.
Shiloh Y (2003) ATM and related protein kinases: safeguarding genome integrity. Nat Rev
Cancer 3(3): 155–168.
Shuen AY & Foulkes WD (2011) Inherited mutations in breast cancer genes--risk and
response. J Mammary Gland Biol Neoplasia 16(1): 3–15.
Smid M, Wang Y, Zhang Y, Sieuwerts AM, Yu J, Klijn JG, Foekens JA & Martens JW
(2008) Subtypes of breast cancer show preferential site of relapse. Cancer Res 68(9):
3108–3114.
Smith HF, Roberts MA, Nguyen HQ, Peterson M, Hartl TA, Wang XJ, Klebba JE, Rogers
GC & Bosco G (2013) Maintenance of interphase chromosome compaction and
homolog pairing in Drosophila is regulated by the condensin cap-h2 and its partner
Mrg15. Genetics 195(1): 127–146.
Smith P, McGuffog L, Easton DF, Mann GJ, Pupo GM, Newman B, Chenevix-Trench G,
kConFab Investigators, Szabo C, Southey M, Renard H, Odefrey F, Lynch H, StoppaLyonnet D, Couch F, Hopper JL, Giles GG, McCredie MR, Buys S, Andrulis I, Senie
R, BCFS BCG, Goldgar DE, Oldenburg R, Kroeze-Jansema K, Kraan J, MeijersHeijboer H, Klijn JG, van Asperen C, van Leeuwen I, Vasen HF, Cornelisse CJ,
Devilee P, Baskcomb L, Seal S, Barfoot R, Mangion J, Hall A, Edkins S, Rapley E,
Wooster R, Chang-Claude J, Eccles D, Evans DG, Futreal P, Nathanson KL, Weber
BL, Breast Cancer Susceptibility Collaboration (UK), Rahman N & Stratton MR
(2006) A genome wide linkage search for breast cancer susceptibility genes. Genes
Chromosomes Cancer 45(7): 646–655.
Solyom S, Aressy B, Pylkäs K, Patterson-Fortin J, Hartikainen JM, Kallioniemi A,
Kauppila S, Nikkilä J, Kosma VM, Mannermaa A, Greenberg RA & Winqvist R
(2012) Breast cancer-associated Abraxas mutation disrupts nuclear localization and
DNA damage response functions. Sci Transl Med 4(122): 122ra23.
Solyom S, Winqvist R, Nikkilä J, Rapakko K, Hirvikoski P, Kokkonen H & Pylkäs K
(2011) Screening for large genomic rearrangements in the FANCA gene reveals
extensive deletion in a Finnish breast cancer family. Cancer Lett 302(2): 113–118.
Somyajit K, Subramanya S & Nagaraju G (2010) RAD51C: a novel cancer susceptibility
gene is linked to Fanconi anemia and breast cancer. Carcinogenesis 31(12): 2031–
2038.
108
Sorensen SA, Mulvihill JJ & Nielsen A (1986) Long-term follow-up of von
Recklinghausen neurofibromatosis. Survival and malignant neoplasms. N Engl J Med
314(16): 1010–1015.
Sorino C, Bruno T, Desantis A, Di Certo MG, Iezzi S, De Nicola F, Catena V, Floridi A,
Chessa L, Passananti C, Cundari E & Fanciulli M (2013) Centrosomal Che-1 Protein
Is Involved in the Regulation of Mitosis and DNA Damage Response by Mediating
Pericentrin (PCNT)-dependent Chk1 Protein Localization. J Biol Chem 288(32):
23348–23357.
Sorlie T, Perou CM, Tibshirani R, Aas T, Geisler S, Johnsen H, Hastie T, Eisen MB, van
de Rijn M, Jeffrey SS, Thorsen T, Quist H, Matese JC, Brown PO, Botstein D,
Lonning PE & Borresen-Dale AL (2001) Gene expression patterns of breast
carcinomas distinguish tumor subclasses with clinical implications. Proc Natl Acad
Sci USA 98(19): 10869–10874.
Soulier J (2011) Fanconi anemia. Hematology Am Soc Hematol Educ Program 2011: 492–
497.
Southey MC, Teo ZL, Dowty JG, Odefrey FA, Park DJ, Tischkowitz M, Sabbaghian N,
Apicella C, Byrnes GB, Winship I, Cancer Family Registry B, Confab K, Baglietto L,
Giles GG, Goldgar D, Foulkes WD & Hopper JL (2010) A PALB2 mutation
associated with high risk of breast cancer. Breast Cancer Res 12(6): R109.
Southey MC, Teo ZL & Winship I (2013) PALB2 and breast cancer: ready for clinical
translation! Appl Clin Genet 6: 43–52.
Srivastava S, Zou ZQ, Pirollo K, Blattner W & Chang EH (1990) Germ-line transmission
of a mutated p53 gene in a cancer-prone family with Li-Fraumeni syndrome. Nature
348(6303): 747–749.
Stacey SN, Manolescu A, Sulem P, Rafnar T, Gudmundsson J, Gudjonsson SA, Masson G,
Jakobsdottir M, Thorlacius S, Helgason A, Aben KK, Strobbe LJ, Albers-Akkers MT,
Swinkels DW, Henderson BE, Kolonel LN, Le Marchand L, Millastre E, Andres R,
Godino J, Garcia-Prats MD, Polo E, Tres A, Mouy M, Saemundsdottir J, Backman
VM, Gudmundsson L, Kristjansson K, Bergthorsson JT, Kostic J, Frigge ML, Geller
F, Gudbjartsson D, Sigurdsson H, Jonsdottir T, Hrafnkelsson J, Johannsson J,
Sveinsson T, Myrdal G, Grimsson HN, Jonsson T, von Holst S, Werelius B, Margolin
S, Lindblom A, Mayordomo JI, Haiman CA, Kiemeney LA, Johannsson OT, Gulcher
JR, Thorsteinsdottir U, Kong A & Stefansson K (2007) Common variants on
chromosomes 2q35 and 16q12 confer susceptibility to estrogen receptor-positive
breast cancer. Nat Genet 39(7): 865–869.
Starink TM, van der Veen JP, Arwert F, de Waal LP, de Lange GG, Gille JJ & Eriksson
AW (1986) The Cowden syndrome: a clinical and genetic study in 21 patients. Clin
Genet 29(3): 222–233.
Stecklein SR & Jensen RA (2012) Identifying and exploiting defects in the Fanconi
anemia/BRCA pathway in oncology. Transl Res 160(3): 178–197.
109
Steffen J, Nowakowska D, Niwinska A, Czapczak D, Kluska A, Piatkowska M,
Wisniewska A & Paszko Z (2006) Germline mutations 657del5 of the NBS1 gene
contribute significantly to the incidence of breast cancer in Central Poland. Int J
Cancer 119(2): 472–475.
Stehelin D, Varmus HE, Bishop JM & Vogt PK (1976) DNA related to the transforming
gene(s) of avian sarcoma viruses is present in normal avian DNA. Nature 260(5547):
170–173.
Stoepker C, Hain K, Schuster B, Hilhorst-Hofstee Y, Rooimans MA, Steltenpool J, Oostra
AB, Eirich K, Korthof ET, Nieuwint AW, Jaspers NG, Bettecken T, Joenje H,
Schindler D, Rouse J & de Winter JP (2011) SLX4, a coordinator of structure-specific
endonucleases, is mutated in a new Fanconi anemia subtype. Nat Genet 43(2): 138–
141.
Stratton MR, Campbell PJ & Futreal PA (2009) The cancer genome. Nature 458(7239):
719–724.
Stratton MR & Rahman N (2008) The emerging landscape of breast cancer susceptibility.
Nat Genet 40(1): 17–22.
Swift M, Sholman L, Perry M & Chase C (1976) Malignant neoplasms in the families of
patients with ataxia-telangiectasia. Cancer Res 36(1): 209–215.
Sy SM, Huen MS & Chen J (2009a) MRG15 is a novel PALB2-interacting factor involved
in homologous recombination. J Biol Chem 284(32): 21127–21131.
Sy SM, Huen MS & Chen J (2009b) PALB2 is an integral component of the BRCA
complex required for homologous recombination repair. Proc Natl Acad Sci USA
106(17): 7155–7160.
Sy SM, Huen MS, Zhu Y & Chen J (2009c) PALB2 regulates recombinational repair
through chromatin association and oligomerization. J Biol Chem 284(27): 18302–
18310.
Syrjäkoski K, Kuukasjärvi T, Waltering K, Haraldsson K, Auvinen A, Borg A, Kainu T,
Kallioniemi OP & Koivisto PA (2004) BRCA2 mutations in 154 finnish male breast
cancer patients. Neoplasia 6(5): 541–545.
Syrjäkoski K, Vahteristo P, Eerola H, Tamminen A, Kivinummi K, Sarantaus L, Holli K,
Blomqvist C, Kallioniemi OP, Kainu T & Nevanlinna H (2000) Population-based
study of BRCA1 and BRCA2 mutations in 1035 unselected Finnish breast cancer
patients. J Natl Cancer Inst 92(18): 1529–1531.
Takashi Y, Kobayashi Y, Tanaka K & Tamura K (2009) Arabidopsis replication protein A
70a is required for DNA damage response and telomere length homeostasis. Plant Cell
Physiol 50(11): 1965–1976.
Taniguchi T, Garcia-Higuera I, Andreassen PR, Gregory RC, Grompe M & D'Andrea AD
(2002) S-phase-specific interaction of the Fanconi anemia protein, FANCD2, with
BRCA1 and RAD51. Blood 100(7): 2414–2420.
Thomas T, Voss AK, Petrou P & Gruss P (2000) The murine gene, Traube, is essential for
the growth of preimplantation embryos. Dev Biol 227(2): 324–342.
110
Thompson D, Duedal S, Kirner J, McGuffog L, Last J, Reiman A, Byrd P, Taylor M &
Easton DF (2005) Cancer risks and mortality in heterozygous ATM mutation carriers.
J Natl Cancer Inst 97(11): 813–822.
Thompson D & Easton D (2004) The genetic epidemiology of breast cancer genes. J
Mammary Gland Biol Neoplasia 9(3): 221–236.
Thompson ER, Boyle SE, Johnson J, Ryland GL, Sawyer S, Choong DY, kConFab,
Chenevix-Trench G, Trainer AH, Lindeman GJ, Mitchell G, James PA & Campbell
IG (2012) Analysis of RAD51C germline mutations in high-risk breast and ovarian
cancer families and ovarian cancer patients. Hum Mutat 33(1): 95–99.
Thull DL & Vogel VG (2004) Recognition and management of hereditary breast cancer
syndromes. Oncologist 9(1): 13–24.
Tischkowitz M & Xia B (2010) PALB2/FANCN: recombining cancer and Fanconi anemia.
Cancer Res 70(19): 7353–7359.
Tominaga K, Kirtane B, Jackson JG, Ikeno Y, Ikeda T, Hawks C, Smith JR, Matzuk MM
& Pereira-Smith OM (2005) MRG15 regulates embryonic development and cell
proliferation. Mol Cell Biol 25(8): 2924–2937.
Tominaga K & Pereira-Smith OM (2002) The genomic organization, promoter position
and expression profile of the mouse MRG15 gene. Gene 294(1–2): 215–224.
Trainer AH, Thompson E & James PA (2011) BRCA and beyond: a genome-first approach
to familial breast cancer risk assessment. Discov Med 12(66): 433–443.
Tuma RS (2009) PARP inhibitors: will the new class of drugs match the hype? J Natl
Cancer Inst 101(18): 1230–1232.
Turnbull C, Ahmed S, Morrison J, Pernet D, Renwick A, Maranian M, Seal S, Ghoussaini
M, Hines S, Healey CS, Hughes D, Warren-Perry M, Tapper W, Eccles D, Evans DG,
Breast Cancer Susceptibility Collaboration (UK), Hooning M, Schutte M, van den
Ouweland A, Houlston R, Ross G, Langford C, Pharoah PD, Stratton MR, Dunning
AM, Rahman N & Easton DF (2010) Genome-wide association study identifies five
new breast cancer susceptibility loci. Nat Genet 42(6): 504–507.
Turnbull C & Rahman N (2008) Genetic predisposition to breast cancer: past, present, and
future. Annu Rev Genomics Hum Genet 9: 321–345.
Turner N, Tutt A & Ashworth A (2005) Targeting the DNA repair defect of BRCA
tumours. Curr Opin Pharmacol 5(4): 388–393.
Tutt A & Ashworth A (2002) The relationship between the roles of BRCA genes in DNA
repair and cancer predisposition. Trends Mol Med 8(12): 571–576.
Tutt A, Robson M, Garber JE, Domchek SM, Audeh MW, Weitzel JN, Friedlander M,
Arun B, Loman N, Schmutzler RK, Wardley A, Mitchell G, Earl H, Wickens M &
Carmichael J (2010) Oral poly(ADP-ribose) polymerase inhibitor olaparib in patients
with BRCA1 or BRCA2 mutations and advanced breast cancer: a proof-of-concept
trial. Lancet 376(9737): 235–244.
UICC Internationa Union Against Cancer. URI: http://www.uicc.org/resources/tnm.
111
Vahteristo P, Bartkova J, Eerola H, Syrjäkoski K, Ojala S, Kilpivaara O, Tamminen A,
Kononen J, Aittomäki K, Heikkilä P, Holli K, Blomqvist C, Bartek J, Kallioniemi OP
& Nevanlinna H (2002) A CHEK2 genetic variant contributing to a substantial
fraction of familial breast cancer. Am J Hum Genet 71(2): 432–438.
Vahteristo P, Eerola H, Tamminen A, Blomqvist C & Nevanlinna H (2001) A probability
model for predicting BRCA1 and BRCA2 mutations in breast and breast-ovarian
cancer families. Br J Cancer 84(5): 704–708.
van der Groep P, van der Wall E & van Diest PJ (2011) Pathology of hereditary breast
cancer. Cell Oncol (Dordr) 34(2): 71–88.
Vaz F, Hanenberg H, Schuster B, Barker K, Wiek C, Erven V, Neveling K, Endt D,
Kesterton I, Autore F, Fraternali F, Freund M, Hartmann L, Grimwade D, Roberts RG,
Schaal H, Mohammed S, Rahman N, Schindler D & Mathew CG (2010) Mutation of
the RAD51C gene in a Fanconi anemia-like disorder. Nat Genet 42(5): 406–409.
Villarroel MC, Rajeshkumar NV, Garrido-Laguna I, De Jesus-Acosta A, Jones S, Maitra A,
Hruban RH, Eshleman JR, Klein A, Laheru D, Donehower R & Hidalgo M (2011)
Personalizing cancer treatment in the age of global genomic analyses: PALB2 gene
mutations and the response to DNA damaging agents in pancreatic cancer. Mol
Cancer Ther 10(1): 3–8.
Vogelstein B & Kinzler KW (1993) The multistep nature of cancer. Trends Genet 9(4):
138–141.
Vogelstein B, Papadopoulos N, Velculescu VE, Zhou S, Diaz LA & Kinzler KW (2013)
Cancer Genome Landscapes. Science 339(6127): 1546–1558.
Vuorela M, Pylkäs K, Hartikainen JM, Sundfeldt K, Lindblom A, von Wachenfeldt
Wappling A, Haanpää M, Puistola U, Rosengren A, Anttila M, Kosma VM,
Mannermaa A & Winqvist R (2011) Further evidence for the contribution of the
RAD51C gene in hereditary breast and ovarian cancer susceptibility. Breast Cancer
Res Treat 130(3): 1003–1010.
Waisfisz Q, de Winter JP, Kruyt FA, de Groot J, van der Weel L, Dijkmans LM, Zhi Y,
Arwert F, Scheper RJ, Youssoufian H, Hoatlin ME & Joenje H (1999) A physical
complex of the Fanconi anemia proteins FANCG/XRCC9 and FANCA. Proc Natl
Acad Sci USA 96(18): 10320–10325.
Walsh T, Casadei S, Lee MK, Pennil CC, Nord AS, Thornton AM, Roeb W, Agnew KJ,
Stray SM, Wickramanayake A, Norquist B, Pennington KP, Garcia RL, King MC &
Swisher EM (2011) Mutations in 12 genes for inherited ovarian, fallopian tube, and
peritoneal carcinoma identified by massively parallel sequencing. Proc Natl Acad Sci
USA 108(44): 18032–18037.
Walsh T & King MC (2007) Ten genes for inherited breast cancer. Cancer Cell 11(2):
103–105.
Wang B (2012) BRCA1 tumor suppressor network: focusing on its tail. Cell Biosci 2(6).
Weinberg RA (1991) Tumor suppressor genes. Science 254(5035): 1138–1146.
Weischer M, Bojesen SE, Ellervik C, Tybjaerg-Hansen A & Nordestgaard BG (2008)
CHEK2*1100delC genotyping for clinical assessment of breast cancer risk: metaanalyses of 26,000 patient cases and 27,000 controls. J Clin Oncol 26(4): 542–548.
112
Welcsh PL, Owens KN & King MC (2000) Insights into the functions of BRCA1 and
BRCA2. Trends Genet 16(2): 69–74.
West SC (2003) Molecular views of recombination proteins and their control. Nat Rev Mol
Cell Biol 4(6): 435–445.
Wiechmann L, Sampson M, Stempel M, Jacks LM, Patil SM, King T & Morrow M (2009)
Presenting features of breast cancer differ by molecular subtype. Ann Surg Oncol
16(10): 2705–2710.
Wijker M, Morgan NV, Herterich S, van Berkel CG, Tipping AJ, Gross HJ, Gille JJ, Pals
G, Savino M, Altay C, Mohan S, Dokal I, Cavenagh J, Marsh J, van Weel M, Ortega
JJ, Schuler D, Samochatova E, Karwacki M, Bekassy AN, Abecasis M, Ebell W,
Kwee ML, de Ravel T & CG M (1999) Heterogeneous spectrum of mutations in the
Fanconi anaemia group A gene. Eur J Hum Genet 7(1): 52–59.
Wilkie AO (1994) The molecular basis of genetic dominance. J Med Genet 31(2): 89–98.
Wooster R, Bignell G, Lancaster J, Swift S, Seal S, Mangion J, Collins N, Gregory S,
Gumbs C & Micklem G (1995) Identification of the breast cancer susceptibility gene
BRCA2. Nature 378(6559): 789–792.
Wooster R, Neuhausen SL, Mangion J, Quirk Y, Ford D, Collins N, Nguyen K, Seal S,
Tran T & Averill D (1994) Localization of a breast cancer susceptibility gene,
BRCA2, to chromosome 13q12–13. Science 265(5181): 2088–2090.
Wu X, Webster SR & Chen J (2001) Characterization of tumor-associated Chk2 mutations.
J Biol Chem 276(4): 2971–2974.
Xia B, Dorsman JC, Ameziane N, de Vries Y, Rooimans MA, Sheng Q, Pals G, Errami A,
Gluckman E, Llera J, Wang W, Livingston DM, Joenje H & de Winter JP (2007)
Fanconi anemia is associated with a defect in the BRCA2 partner PALB2. Nat Genet
39(2): 159–161.
Xia B, Sheng Q, Nakanishi K, Ohashi A, Wu J, Christ N, Liu X, Jasin M, Couch FJ &
Livingston DM (2006) Control of BRCA2 cellular and clinical functions by a nuclear
partner, PALB2. Mol Cell 22(6): 719–729.
Yaghjyan L & Colditz GA (2011) Estrogens in the breast tissue: a systematic review.
Cancer Causes Control 22(4): 529–540.
Yamamoto K, Ishiai M, Matsushita N, Arakawa H, Lamerdin JE, Buerstedde JM,
Tanimoto M, Harada M, Thompson LH & Takata M (2003) Fanconi anemia FANCG
protein in mitigating radiation- and enzyme-induced DNA double-strand breaks by
homologous recombination in vertebrate cells. Mol Cell Biol 23(15): 5421–5430.
Yang H, Jeffrey PD, Miller J, Kinnucan E, Sun Y, Thoma NH, Zheng N, Chen PL, Lee
WH & Pavletich NP (2002) BRCA2 function in DNA binding and recombination
from a BRCA2-DSS1-ssDNA structure. Science 297(5588): 1837–1848.
Yang H, Li Q, Fan J, Holloman WK & Pavletich NP (2005) The BRCA2 homologue Brh2
nucleates RAD51 filament formation at a dsDNA-ssDNA junction. Nature 433(7026):
653–657.
Yoshida K & Miki Y (2004) Role of BRCA1 and BRCA2 as regulators of DNA repair,
transcription, and cell cycle in response to DNA damage. Cancer Sci 95(11): 866–871.
113
Yu DS, Sonoda E, Takeda S, Huang CL, Pellegrini L, Blundell TL & Venkitaraman AR
(2003) Dynamic control of Rad51 recombinase by self-association and interaction
with BRCA2. Mol Cell 12(4): 1029–1041.
Yuan SS, Lee SY, Chen G, Song M, Tomlinson GE & Lee EY (1999) BRCA2 is required
for ionizing radiation-induced assembly of Rad51 complex in vivo. Cancer Res 59(15):
3547–3551.
Zemunik T & Boraska V (2011) Genetics of Type 1 Diabetes. In: Wagner D (ed) Type 1
Diabetes - Pathogenesis, Genetics and Immunotherapy. InTech.
Zhang F, Fan Q, Ren K & Andreassen PR (2009a) PALB2 functionally connects the breast
cancer susceptibility proteins BRCA1 and BRCA2. Mol Cancer Res 7(7): 1110–1118.
Zhang F, Ma J, Wu J, Ye L, Cai H, Xia B & Yu X (2009b) PALB2 links BRCA1 and
BRCA2 in the DNA-damage response. Curr Biol 19(6): 524–529.
Zhang P, Du J, Sun B, Dong X, Xu G, Zhou J, Huang Q, Liu Q, Hao Q & Ding J (2006)
Structure of human MRG15 chromo domain and its binding to Lys36-methylated
histone H3. Nucleic Acids Res 34(22): 6621–6628.
Zhang S, Hemmerich P & Grosse F (2007) Centrosomal localization of DNA damage
checkpoint proteins. J Cell Biochem 101(2): 451–465.
Zheng Y, Zhang J, Hope K, Niu Q, Huo D & Olopade OI (2010) Screening RAD51C
nucleotide alterations in patients with a family history of breast and ovarian cancer.
Breast Cancer Res Treat 124(3): 857–861.
114
Original publications
I
Haanpää M, Reinman M, Nikkilä J, Erkko H, Pylkäs K & Winqvist R (2009)
Mutation analysis of the AATF gene in breast cancer families. BMC Cancer 9(1): 457.
II Rio Frio T, Haanpää M, Pouchet C, Pylkäs K, Vuorela M, Tischkowitz M, Winqvist R
& Foulkes WD (2010) Mutation analysis of the gene encoding the PALB2-binding
protein MRG15 in BRCA1/2-negative breast cancer families. J Hum Genet 55(12):
842–843.
III Haanpää M, Pylkäs K, Moilanen JS & Winqvist R (2013) Evaluation of the need for
routine clinical testing of PALB2 c.1592delT mutation in BRCA negative Northern
Finnish breast cancer families. BMC Med Genet 14(1): 82.
IV Mantere T, Haanpää M, Hanenberg H, Schleutker J, Kallioniemi A, Kähkönen M,
Parto K, Avela K, Aittomäki K, von Koskull H, Hartikainen J, Kosma VM, Laasanen
SL, Mannermaa A, Pylkäs K & Winqvist R (2014) Finnish Fanconi anemia mutations
and hereditary predisposition to breast and prostate cancer. Manuscript.
These articles have been reprinted with the kind permission of their copyright
holders.
Original publications are not included in the electronic version of the dissertation.
115
116
ACTA UNIVERSITATIS OULUENSIS
SERIES D MEDICA
1230. Vaaramo, Kalle (2013) Alcohol affects the outcome after head trauma
1231. Moilanen, Riina (2013) Heterotopic ossification in skin : special focus on multiple
miliary osteoma cutis and the role of bone morphogenetic proteins
1232. Vatjus, Ritva (2014) Kohti suhdekeskeisyyttä lääkärin ja potilaan kohtaamisessa :
laadullinen tutkimus potilas-lääkärisuhteen hahmottumisesta yleislääkäreiden
koulutuksessa
1233. Mankinen, Katariina (2014) Neuropsychological performance and functional MRI
findings in children with non-lesional temporal lobe epilepsy
1234. Karvonen, Henna (2014) Ultrastructural and functional characterization of
myofibroblasts in lung diseases
1235. Sipola, Seija (2014) Colectomy in an ICU patient population : clinical and
histological evaluation
1236. Lipponen, Kaija (2014) Potilasohjauksen toimintaedellytykset
1237. Jansson, Miia (2014) The effectiveness of education on critical care nurses'
knowledge and skills in adhering to guidelines to prevent ventilator-associated
pneumonia
1238. Turunen, Sanna (2014) Protein-bound citrulline and homocitrulline in rheumatoid
arthritis : confounding features arising from structural homology
1239. Kuorilehto, Ritva (2014) Moniasiantuntijuus sosiaali- ja terveydenhuollon
perhetyössä : monitahoarviointi Q-metodologialla
1240. Sova, Henri (2014) Oxidative stress in breast and gynaecological carcinogenesis
1241. Tiirinki, Hanna (2014) Näkyvien ja piilotettujen merkitysten rajapinnoilla :
terveyskeskukseen liittyvät kulttuurimallit asiakkaan näkökulmasta
1242. Syrjänen, Riikka (2014) TIM family molecules in hematopoiesis
1243. Kauppila, Joonas (2014) Toll-like receptor 9 in alimentary tract cancers
1244. Honkavuori-Toivola, Maria (2014) The prognostic role of matrix
metalloproteinase-2 and -9 and their tissue inhibitor-1 and -2 in endometrial
carcinoma
1245. Pienimäki, Tuula (2014) Factors, complications and health-related quality of life
associated with diabetes mellitus developed after midlife in men
1246. Kallio, Miki (2014) Muuttuuko lääketieteen opiskelijoiden käsitys terveydestä
peruskoulutuksen aikana : kuusivuotinen seurantatutkimus
Book orders:
Granum: Virtual book store
http://granum.uta.fi/granum/
D 1247
OULU 2014
UNIV ER S IT Y OF OULU P. O. B R[ 00 FI-90014 UNIVERSITY OF OULU FINLAND
U N I V E R S I TAT I S
S E R I E S
SCIENTIAE RERUM NATURALIUM
Professor Esa Hohtola
HUMANIORA
University Lecturer Santeri Palviainen
TECHNICA
Postdoctoral research fellow Sanna Taskila
MEDICA
Professor Olli Vuolteenaho
SCIENTIAE RERUM SOCIALIUM
ACTA
U N I V E R S I T AT I S O U L U E N S I S
Maria Haanpää
E D I T O R S
Maria Haanpää
A
B
C
D
E
F
G
O U L U E N S I S
ACTA
A C TA
D 1247
HEREDITARY
PREDISPOSITION TO BREAST
CANCER – WITH A FOCUS
ON AATF, MRG15, PALB2, AND
THREE FANCONI ANAEMIA
GENES
University Lecturer Veli-Matti Ulvinen
SCRIPTA ACADEMICA
Director Sinikka Eskelinen
OECONOMICA
Professor Jari Juga
EDITOR IN CHIEF
Professor Olli Vuolteenaho
PUBLICATIONS EDITOR
Publications Editor Kirsti Nurkkala
ISBN 978-952-62-0457-4 (Paperback)
ISBN 978-952-62-0458-1 (PDF)
ISSN 0355-3221 (Print)
ISSN 1796-2234 (Online)
UNIVERSITY OF OULU GRADUATE SCHOOL;
UNIVERSITY OF OULU, FACULTY OF MEDICINE,
INSTITUTE OF CLINICAL MEDICINE, DEPARTMENT OF CLINICAL GENETICS;
FACULTY OF MEDICINE, INSTITUTE OF DIAGNOSTICS,
DEPARTMENT OF CLINICAL CHEMISTRY;
BIOCENTER OULU;
OULU UNIVERSITY HOSPITAL;
NATIONAL GRADUATE SCHOOL OF CLINICAL INVESTIGATION
D
MEDICA