Sei sulla pagina 1di 10

doi: 10.1111/j.1365-2796.2012.02525.

x
Prevention of breast cancer in the context of a national breast
screening programme
A. Howell
1,2
, S. Astley
2
, J. Warwick
3
, P. Stavrinos
1
, S. Sahin
1
, S. Ingham
4
, H. McBurney
5
, B. Eckersley
1
, M. Harvie
1
,
M. Wilson
1
, U. Beetles
1
, R. Warren
6
, A. Hufton
2
, J. Sergeant
2
, W. Newman
5
, I. Buchan
4
, J. Cuzick
3
&D. G. Evans
1,5
Fromthe
1
GenesisPreventionCentreandNightingaleBreast ScreeningCentre, UniversityHospital of SouthManchester;
2
School of Cancer and
EnablingSciences, Universityof Manchester, Manchester;
3
Centrefor Cancer Prevention, WolfsonInstituteof PreventiveMedicine, QueenMary
Universityof London, London;
4
School of CommunityBasedMedicine, Universityof Manchester, Manchester;
5
Genetic Medicine, Manchester
AcademicHealthSciencesCentre, Universityof Manchester andCentral Manchester FoundationTrust, Manchester; and
6
CambridgeBreast Unit,
AddenbrookesHospital, Cambridge; UK
Abstract. Howell A, Astley S, Warwick J, Stavrinos P,
SahinS, InghamS, McBurney H, Eckersley B, Harvie
M, WilsonM, BeetlesU, WarrenR, HuftonA, Sergeant
J, Newman W, Buchan I, Cuzick J, Evans DG
(Genesis Prevention Centre and Nightingale Breast
Screening Centre, University Hospital of South Man-
chester; School of Cancer and Enabling Sciences,
University of Manchester, Manchester; Centre for
Cancer Prevention, Wolfson Institute of Preventive
Medicine, QueenMaryUniversityof London, London;
School of Community Based Medicine, University of
Manchester, Manchester; Genetic Medicine, Man-
chester Academic HealthSciences Centre, University
of Manchester and Central Manchester Foundation
Trust, Manchester; and Cambridge Breast Unit,
Addenbrookes Hospital, Cambridge; UK). Prevention
of breast cancer in the context of a national breast
screening programme (Review). J Intern Med 2012;
271: 321330.
Breast cancer is not only increasing in the West but
also particularly rapidly in Eastern countries where
traditionally the incidence has been low. The rise in
incidenceismainlyrelatedtochangesinreproductive
patterns and lifestyle. These trends could potentially
be reversed by dening women at greatest risk and
offeringappropriatepreventivemeasures. Amodel for
this approach was the establishment of Family His-
tory Clinics (FHCs), which have resulted in improved
survival in younger women at high risk. New predic-
tive models of risk that include reproductive and
lifestyle factors, mammographic density and mea-
surement of risk-associated single nucleotide poly-
morphisms(SNPs) maygivemorepreciseinformation
concerning risk and enable better targeting for mam-
mographic screening programmes and of preventive
measures. Endocrine prevention using anti-oestro-
gens andaromatase inhibitors is effective, andobser-
vational studies suggest lifestyle modication may
alsobeeffective. However, referral toFHCsisopportu-
nisticandpredominantlyincludesyounger women. A
better approach for identifying older women at risk
maybetousenational breast screeningprogrammes.
Here were described pilot studies to assess whether
the routine assessment of breast cancer risk is feasi-
ble withina population-basedscreening programme,
whether the feedback and advice on risk-reducing
interventions would be welcomed and taken up, and
to consider whether the screening interval should be
modiedaccordingtobreast cancerrisk.
Keywords: breast cancer, prevention, risk estimation,
screening.
Introduction
Breast cancer is the most common female cancer in
many countries. In the UK, the age-adjusted rate for
breast cancer rose from 80 to 100 per 100 000 wo-
menbetween 1985 and 2007 [1]. Breast cancer rates
are rising particularly rapidly in countries with a tra-
ditionallylowincidencesuchasAfrica, Asiaandparts
of SouthAmerica[2]. InIceland, long-termprecisere-
cord keeping has enabled accurate comparisons of
breast cancer incidence between 19211944 and
19852002. During this time period, there was a
fourfold increase, not only in the incidence of spo-
radic breast cancer, but also of the disease incarriers
of the single founder Icelandic BRCA2mutation(pop-
ulationincidence 0.5), indicating that environmental
factors not only affect risk in the general population
but alsoaffect mutationpenetrance[3].
The marked increases in breast cancer incidence are
believed to be mainly related to changes in reproduc-
tive and lifestyle factors and the use of hormone
2012 The Association for the Publication of the Journal of Internal Medicine 321
Review
|
replacement therapy (HRT). It is estimated that our
female ancestors had about 160 ovulations over
their reproductive period whereas recently it was
estimated that women nowhave over 450 ovulations
during reproductive life; thus, the risk of breast can-
cer presumed to be associated with cyclical activity
of breast epithelial cell proliferation during the men-
strual cycle is increased [4]. Lifestyle-related breast
cancer risk factors include obesity and lack of physi-
cal activity, both of which are becoming increasingly
prevalent. Huang et al. [5] reported in the Nurses
Health Study that breast cancer risk doubled in
women who gained 20 kg or more in weight between
the ages of 20 and 50 years compared to women who
were essentially weight stable; many other studies
report similar relative risks. A recent overview of the
contribution of body mass index (BMI) to the inci-
dence of breast cancer estimated that that risk
increased approximately 14%for each5 kg m
)2
gain
weight in this population [6]. Compared with no
physical activity, more than 4 h per week is esti-
mated to reduce the risk of breast cancer by approxi-
mately 25% [7]. Thus, although it is not feasible to
change reproductive patterns, risks may neverthe-
less be reduced by avoiding the use of HRT, blocking
the tumour-promoting effects of endogenous
hormones with agents such as tamoxifen and
raloxifene, and sustained lifestyle changes will also
probably be effective. Although we know that such
measures are likely to be effective, it is not clear how
best to identify a target population or whether risk-
reducing interventions would be acceptable and
sustainable within the population. Preventive
approaches have been undertaken for some years in
the context of clinics for younger women at risk,
mainly because of a family history, and are a para-
digm for targeting the much larger population of
older women.
The Family History Clinic model
Family History Clinics (FHCs) were rst set up in the
1980s in response to womens increasing awareness
of their geneticriskof breast cancer [8, 9]. Clinicsgen-
erally offer risk information, mammographic and
other screening and advice concerning the appropri-
ateness of particular preventive interventions for wo-
men[8, 9]. Informationisgivenconcerningtheproba-
bility of carrying a mutationina cancer gene, offering
genetic testing in the family if appropriate [1014]
andassessingoverall riskof breast cancer for noncar-
riers produced by combining genetic and other risk
factorsbytheuseof, for example, modelssuchasGail
et al. andTyrer et al. [15, 16].
It is customary to offer annual screening by mam-
mography in this group from age 35 or 40 although
the effectiveness of this as regards reducing breast
cancer mortalityhas never beenassessedinrandom-
ized controlled trials. As a surrogate, we compared
the survival of women undergoing 1218 monthly
mammographyinour FHCandwhodevelopedbreast
cancer with patients in the same age range who pre-
sented symptomatically our surgical clinic over a 10-
year period. After correction for lead time bias, sur-
vival [HR 0.24 (95% CI 0.080.43) P = 0.005] and
disease-free survival [HR 0.25 (95% CI 0.110.57)
P < 0.001] were signicantly improved in the FHC
compared with women who presented symptomati-
cally to the same breast unit [17]. More recently, in a
multicentre study in76FHCs inthe UK, the outcome
of 6710 women who had annual mammograms aged
4049 were compared with 106 971 women aged be-
tween 40and42fromthe general populationat aver-
age riskwho were followedthroughtheir forties inthe
UK Age Trial and a Dutch study (cancer cases be-
tween25and77 years of womenwithafamilyhistory
of breast cancer). Compared with the two control
groups, screened women were more likely to have
small node negative tumours. The predicted 10-year
deaths frombreast cancer was 1.1%in the FHCcon-
trols compared with 1.38% in the controls from the
UK Age Trial. The relative risk reduction was 0.80
(95%CI 0.660.96, P = 0.022) and suggests that an-
nual mammography in FHCs is likely to prevent
deaths from breast cancer [18]. Other studies in
BRCA1 and BRCA2 carriers indicate that risk-reduc-
ing oophorectomy and mastectomy are also associ-
ated with improved survival compared with no sur-
gery groups [18, 19], although recent modelling [20]
indicates that focussed screening by mammography
and magnetic resonance imaging (MRI) may be as
effective as bilateral prophylactic mastectomy. How-
ever, there must be some doubt about the efcacy of
screeninginBRCA1carriersbecauseevensmall node
negative tumours have a poorer prognosis thanother
screendetectedcancer [21].
Models of risk estimation used in Family History Clinics
Two types of risk are computed in FHCs. One is the
probability of being a mutation carrier, which is
estimated from family history [1014] and the other
the overall risk of developing breast cancer within a
specic time period, which is calculated from fam-
ily history and other established breast cancer risk
factors [15, 16]. For example, the BOADICEA model
developed in Cambridge and the Manchester Score
predict the probability of a mutation based on the
A. Howell et al.
|
Review: Prevention of breast cancer
322 2012 The Association for the Publication of the Journal of Internal Medicine
Journal of Internal Medicine, 2012, 271; 321330
number of breast, ovary and other relevant cancers
in the family to determine the need for mutation
testing [11, 12]. Mutation prediction of these mod-
els have been improved more recently by adding
histological details of diagnosed cancers [13, 14] In
women without mutations, the model developed by
Gail et al. [15] in 1989 and subsequently modied
by Constantino et al. [22] is the most widely used
model in the United States. The Gail model combines
age, number of rst degree relatives with breast
cancer, number of breast biopsies and whether the
biopsies contain epithelial atypia. The model was
developed in three hundred thousand women
undergoing breast screening in the Breast Cancer
Detection Demonstration Project between 1973 and
1980 and validated in the Nurses Health Study
[23]. The modied model focuses on invasive breast
cancer only and is derived from the USA Surveil-
lance Epidemiology and End Results Database and
now includes incidence rates of African American
patients [22].
The TyrerCuzick Risk Prediction Model was devel-
oped to incorporate additional risk factors compared
withGail inanattempt topredict individual riskmore
accurately and to be more relevant to European pop-
ulations [16]. Based on information from the IBIS-1
Breast Cancer Prevention Trial and additional epide-
miological data, themodel includes ageof breast can-
cer ina relative, ovariancancer inthe family, second-
degree relatives with breast and or ovarian cancer,
BMI, age at menopause anduse of HRT inaddition to
factorsusedintheGail 1and2models[15] (Fig. 1).
To choose a model for use in our FHC, we assessed
howwell eachmodel predicted52cancers detectedin
1933 women in our FHC with a mean follow-up of
5.27 years. The ratios (with 95% condence inter-
vals) of expectedtoobservedcancerswere0.48(0.37
0.64) for Gail, 0.56 (0.430.75) for Claus, 0.49(0.37
0.65) for Ford and 0.81 (0.621.08) for TyrerCuzick
[24]. Thus, the Gail model underpredictedinour pop-
ulation as has beenreported by other studies in Eur-
ope and the USA [2527]. In our clinic, we found that
the TyrerCuzick model performs optimally but this
result requires further validation in other high-risk
populations and also in the general population. Fur-
ther discussion of the merits of these and other risk
predictionmodelsmay befoundintworecent reviews
[28, 29].
Addition of mammographic density to risk models
Mammographic breast density is a major risk factor
for breast cancer. Density may be subjectively quan-
tied visually by assessing the proportion of dense
versus fatty tissue and is variously subdivided into
between four and 21 categories (Fig. 2). Two recent
meta-analyses reported a four- to ve-fold increased
risk of invasive cancer for women inthe highest cate-
goryof breast densitycomparedwiththoseinthelow-
est [30, 31] (Fig. 1). Cummings et al. [31] performed a
meta-analysis of how much breast density added to
thediscriminativecapacityof breast cancer riskmod-
els, based on standard risk factors, instudies involv-
ing a total of 12 754 women. The area under (AUC)
the receiver operator characteristic curve (ROC.
Risk: During screening period Lifetime
30
33.0%
35
High risk
2 FDRs
Menarche 10
FFTP 34
25
20
26.4%
Mammographic
density 50%
B
r
e
a
s
t

c
a
n
c
e
r

(
%
)
15
10
5.4%
9%
Population risk
5
0
0 35 47 74 84
1.61% 2.8%
Low risk
0 FDRs
Menarche 14
FFTP20
Follow-up time (years)

Mammographic
density 5%
Fig. 1 Lifetime risk of breast
cancer andriskduringthe period
of screening in women with in-
creased risk factors, population
risk factors and minimal risk
factors.
A. Howell et al.
|
Review: Prevention of breast cancer
2012 The Association for the Publication of the Journal of Internal Medicine 323
Journal of Internal Medicine, 2012, 271; 321330
c-statistic) ranged from 0.62 to 0.66 compared with
thewidelyreportedAUCfor the Gail model of between
0.58 and 0.62. In a study from California, a model
incorporating Gail [32] and additional risk factors re-
portedanAUCof 0.605without densityand0.62with
density (measured visually by the four category BIR-
ADS system). Tice et al. [33] used a simplied Gail
model andreportedthat theAUCincreasedfrom0.61
to0.66alsousingBIRADSdensity. Chenet al. [34] re-
portedimprovedestimatesof absoluteriskwhenden-
sity assessed by the CUMULUS systemof computer-
aideddenseareacalculationwasaddedtotheassess-
ment of the Gail model. Thus, these studies suggest
that combining classical breast cancer risk factors
and mammographic density inan appropriate model
increases the discriminatorypower of riskprediction,
but byarelativelysmall magnitude.
Addition of single nucleotide polymorphismresults to risk models
Recent genome-wide association studies (GWAS)
have identied various novel breast cancer suscepti-
bility variants [3538]. Each single nucleotide poly-
morphisms (SNP) is associated with a small increase
or reductioninrisk, andit ishopedthat bycombining
individual SNP risks and as more SNPs are discov-
ered, the overall estimates of breast cancer risk will
be improved. Van Zitteren et al. [39] performed a
meta-analysis of known breast cancer risk SNPs and
modelled the effect on the AUC of ROC curves in a
simulated population of 10 000 women. They identi-
ed41polymorphisms that were signicantlyassoci-
atedwithbreast cancer riskandcomputedanAUCof
0.67, which is similar to the best risk assessment
model with density as outlined above. Others have
modelledhowmuchsmall numbersof validatedSNPs
add to risk prediction models, mainly using the Gail
model. For example, by adding seven SNPs, Gail re-
ported an increase in the AUC from 0.607 to 0.632
[40]. Other studies using from7 to 11SNPs have also
demonstrated small increases in the AUC compared
with risk models alone [4043]. Park et al. [44] mod-
elled the effect of 67 susceptibility SNPs used alone
andestimatedAUCof 0.635andcommentedthat risk
models based on common variants appear to have
modest discriminatory power. These estimates are
madeassumingnogeneticinteractionbetweenSNPs,
which may or may not be true. If there is no interac-
tion and the fact that there are now a reported addi-
tional 30 new SNPs (see Easton et al. this issue),
there is a potential to use these and other known
SNPs to improve predictionfurther. Amajor question
is whether combining the current standardriskmod-
elswithdensityandalsoSNPswill improveprediction
further withthehopethat moreprecisepredictionwill
allow us to target preventive measures more accu-
rately (See reviews by Chatterjee et al. [45] Predictng
the future of genetic risk prediction, and by Pharaoh
et al. [46]). Ultimately, it is essential that any com-
bined models are validated properly in appropriate
populationsof womenbeforetheir widespreaduse.
Encouragingly, recent studies indicate no interac-
tion between SNPs and other standard risk factors.
Milne et al. [47] found no interaction between 12
SNPs and age at menarche, ever having a live birth,
number of live births, age at rst birth and BMI in
data from 26 349 invasive breast cancer cases and
up to 32 208 controls. In the UK Million Women
Study, Travis et al. [48] reported no interaction be-
tween 12 SNPs and ten established breast cancer
risk factors (age at menarche, parity, age at rst
birth, breastfeeding, menopausal status, age at
menopause, use of HRT, BMI, height and alcohol
consumption) after correction for multiple testing. In
the Predicting Risk Of Cancer At Screening (PRO-
CAS) study (see later), little or no interaction was
found between high-risk women detected by the
Tyer-Cuzick model, mammographic density and 18
SNPs (Evans et al. submitted). Clearly more work is
needed to resolve the paradox that mammographic
density and SNPs appear to add only a little to the
discriminatory power of the Gail model whereas
other studies, including our own, showlittle interac-
tion between standard risk factors, mammographic
density andSNPs.
Fig. 2 Screening interval re-
latedtoriskassumingaconstant
pick up rate of four breast can-
cers per 1000 women screened.
The interval for high risk is esti-
mated at 15 months, for popula-
tion risk is 36 months and for
lowrisk, 70 months.
A. Howell et al.
|
Review: Prevention of breast cancer
324 2012 The Association for the Publication of the Journal of Internal Medicine
Journal of Internal Medicine, 2012, 271; 321330
Breast cancer prevention
The question arises about the effectiveness of cur-
rently available preventive measures and how we
might target the appropriate at risk population more
accurately. Two main preventive avenues of investi-
gation have been undertaken: the use of endocrine
blocking agents (mainly anti-oestrogen) and lifestyle
change, particularlyweight lossandexercise.
Endocrine prevention
Recent reviews summarize the data indicating the
effectiveness of preventive therapy (chemo-preven-
tion) of breast cancer [31, 49, 50]. The agents already
shown to be effective include tamoxifen and raloxif-
ene [5158]. More recently, the aromatase inhibitor
exemestane was reported to give greater risk reduc-
tion than placebo [53]; a further study of anastrozole
versus placebo (IBIS II) is in progress [54]. Most pre-
ventionstudies are performed inwomenat increased
risk of breast cancer. For example, the NSABP1
(tamoxifen v placebo [55]) and STAR (tamoxifen v ra-
loxifene [52]) trials stipulate that only women with a
5-year Gail risk of breast cancer of >1.66%should be
entered into these studies. However, raloxifene was
shown to be effective in women with population
breast cancer risk but who were treated with raloxif-
ene for osteoporosis [TheMOREtrial [56]) andfor car-
diovascular disease(TheRUTHtrial [57]).
Cuzick et al. [51] performed an overview of the four
randomized, placebo-controlled trials of tamoxifen
and reported anoverall risk reduction of 38%[51]. In
the IBIS I trial, the long-term effect of 5 years of
tamoxifen and a further 5 years of follow-up showed
that thecurvesfor placeboandtreatedgroupscontin-
uedto diverge so that there was a doubling of the pre-
ventive effect at 10 years compared with the effect
determined at 5 years [58]. In the STAR trial, women
were randomly assigned to receive either tamoxifen
or raloxifene for 5 years [52]. The risk ratio (RR: ra-
loxifene versus tamoxifen) for invasive breast cancer
was 1.24 (95% CI, 1.051.47). The greater effective-
ness of tamoxifen was associated with a higher inci-
dence of side effects so that the risk benet ratio fa-
voured raloxifene in women with a uterus and was
equivalent to tamoxifen in women without a uterus,
indicating the important negative effect of tamoxifen
ontheendometrium[59].
Aromatase inhibitors are more effective for the pre-
vention of relapse after breast cancer diagnosis com-
pared with tamoxifen, and the early results of a com-
parison between exemestane versus placebo for
prevention of breast cancer in women at high risk
showa 62%reduction inrisk [OR0.38(95%CI)] with
little differences inthe side effect proles betweenex-
emestane and placebo [53]. However, the estimated
number needed to treat (NNT) to prevent one breast
cancer was projected to be 25. These and other NNT
data indicate the need for more precise prediction of
risk, for more effective agents and for biomarkers to
predict women most likely to benet frompreventive
therapy [60]. Currently tamoxifen is the preventive
treatment of choice for premenopausal women and
raloxifene for postmenopausal women. Aromatase
inhibitors, although promising, require further data
ontheir riskbenet ratio.
Lifestyle change
There are few randomized data to support a positive
effect of lifestyle change in relation to breast cancer
prevention. However, observational data indicate
that lifestyle, mainly caloric excess and exercise
deprivation, increases the risk of breast cancer and
that breast cancer risk is reduced by decreasing
weight and increasing physical activity. Two large
prospective studies [61, 62] demonstrate that weight
reduction in midlife or after the menopause de-
crease the risk of postmenopausal breast cancer by
approximately 2550% as does weight reduction re-
lated to bariatric surgery [63]. The results of other
observational studies of weight reduction are mixed,
possibly reecting the small size of some and lack of
data on maintained weight loss in the reported stud-
ies (summarized in [64, 65]). Reduction in fat intake
without appreciable calorie restriction has only a
minor affect on risk as shown in the Womens Health
Initiative large randomized trial [65]. This study also
demonstrated that increased intake of vegetables,
fruit and grain does not appear to reduce breast
cancer risk.
A meta-analysis of 93 studies of exercise and breast
cancer incidence reportedanoverall riskreductionof
about 25%in both pre- and postmenopausal women
[9]. The risk reduction was greatest in women with a
normal BMI, suggesting that the optimal approachto
lifestyle reduction of risk of breast cancer is to com-
bineweight control andappropriatephysical activity.
Use of screening programmes to focus risk reduction
Most Western and many other countries fund mam-
mographic screening programmes designed to detect
breast cancer early with aimof increasing cure rates.
A. Howell et al.
|
Review: Prevention of breast cancer
2012 The Association for the Publication of the Journal of Internal Medicine 325
Journal of Internal Medicine, 2012, 271; 321330
Beyond the system of FHCs that tends to focus on
younger women with a family history, there are few
systematic attempts to offer risk estimation and pre-
ventive measures inthe context of screening [32, 33].
If such an approach is advisable, the question arises
How might such a system be instituted? The rise in
incidence of breast cancer, the trauma of breast can-
cer diagnosis and treatment and indications of the
effectiveness of endocrine and lifestyle prevention
make it appropriate to explore measures to reduce
breast cancer risk by introducing risk predictionand
preventive measures in the context of mammo-
graphicscreeningprogrammes.
Family History Clinics depend upon women present-
ing to their family doctors and being referred if above
a certain risk threshold. The emphasis on family his-
tory tends to focus the service on younger women
ignoring other important risk factors such as age of
rst pregnancy. Most breast cancer is diagnosed in
older womenduringthe screeningperiod[[31, 4772]
in the UK National Health Service Breast Screening
Programme (NHSBSP) Fig. 1]. The majority of older
women do not have a family history but have other
risk factors. Assessing riskat screening is potentially
attractive because a large proportion of the female
population is screened on a regular basis, it is not
opportunistic and would involve a group of older
womenlargely not seen inthe FHCsystem. However,
because screening large numbers of women effec-
tively requires high throughput, it would be vitally
important that adding risk prediction did not inter-
fere with the mechanics of the screening process or
reduce in the numbers of women attending for mam-
mography.
The PROCAS study
We have initiated the PROCAS study to test the feasi-
bility and acceptability to women of collecting breast
cancer risk information (standard risk factors, mam-
mographic density and SNPs) during the routine
mammographicscreeningprocessintheUKNHSBSP.
Theaimof thestudyisattempt toimproveriskpredic-
tion and introduce preventive strategies in women at
high risk. Unlike most other screening programmes,
the interval betweenmammograms inthe NHSBSPis
3 years leading to high rates of interval cancers [66,
67] andthus a longer termaimof PROCASis toinves-
tigate the possibility of adjusting the screening inter-
val basedonrisk.
The study was initiated in October 2009 in fteen
screening sites in Manchester, mainly on mobile
screening vans. By August 2011, 30 000 of the pro-
jected 60 000 women were recruited. Women are
mailed a risk questionnaire (see PROCAS website
http://www.uhsm.nhs.uk/research/Pages/PROCAS
study.aspx) in the interval between the invitation to
screening and attending for a mammogram, because
this was found to be the optimal approach in a previ-
ous large-scale study (CADET [68]). Approximately
70% of women invited attend for screening, and, to
date, 40%of themhave agreed to enter PROCAS. The
risk questionnaire is completed prior to the mam-
mography appointment and returned at the time of
mammography where informed consent is obtained.
The questionnaire is scanned into the database later
and the TyrerCuzick risk automatically computed.
Mammographic density is assessed on a visual ana-
logue scale (VAS) by 11 specially trained radiologists
andradiographers. Threevolumetricmethodsof den-
sity calculation (Volpara, Quantra and Stepwedge)
are being compared with CUMULUS [69] and VAS,
but the ultimate aimis to automate measurement so
that the risk fromdensity can be read out with stan-
dard risk factors automatically [70]. Ten per cent of
the population give buccal smear samples for SNP
estimation(currently18SNPsaretested).
Assessment of Tyer-Cuzick risk and VAS mammo-
graphic density have been reported for the rst
10 000womenandSNPinformationavailableon983
women (Evans et al. submitted). The median 10-year
breast cancer risk was 2.65%, and the median VAS
score was approximately 25%. Interestingly, when
the top 5%of women for TyrerCuzick risk, VAS den-
sity andSNPs were compared, there was little overlap
in the populations identied, suggesting that the
threemethodsdetect different at riskpopulations.
Women with a 10-year risk of 8% or with a 10-year
risk of 57.9%and with a mammographic density of
60% (VAS) were invited to attend or be telephoned
to be counselled concerning their risk in our FHC.
Over 80% have been counselled to date and 18.8%
of 85 eligible women at high risk entered a random-
ized prevention study. Thus, to date, results from
the PROCAS study indicate that it is feasible to as-
sess breast cancer risk and offer risk information
and risk-reducing advice within the context of the
population mammographic screening. The utility
for improved risk calculation by combining stan-
dard risk factors in the TyrerCuzick model, mam-
mographic density and SNP estimations will be as-
sessed after each individuals second mammogram
at 3 years when we expect about 600 tumours in
the population of 60 000 women enrolled. A similar
A. Howell et al.
|
Review: Prevention of breast cancer
326 2012 The Association for the Publication of the Journal of Internal Medicine
Journal of Internal Medicine, 2012, 271; 321330
programme to PROCAS called KARMA involving
100 000 women has been initiated in Sweden by
Hall et al. (personal communication).
Prospects for risk-adapted mammographic screening
Unlike most other screening programmes, the inter-
val betweenmammograms inthe NHSBSPis 3 years.
A recent report of the NHSBSP indicated that 38%of
invasive tumours detected in the context of the pro-
gramme were found inthe interval between mammo-
grams. Interval cancers have a poorer prognosis and
reduce the potential effectiveness of the programme
[64]. Identicationof womenlikelyto developinterval
cancersandofferingthemtailoredscreeningandpre-
ventive interventions may be a way to improve the
effectivenessof theNHSBSP. Thereisevidencetosug-
gest that womenat highriskof breast cancer aremore
likely to develop interval cancers. The Swedish two-
county study showed that women with a family his-
tory of breast cancer were signicantly more likely to
developbreast cancer inthe interval betweenscreens
than equivalent women with no family history [72].
High mammographic breast density also increases
the risk of developing interval breast cancer [71].
Thus, it may be appropriate to offer women at high
risk and with high density a shorter screening inter-
val. Women at very lowrisk of developing breast can-
cer mayrequirescreeninglessfrequentlyandthereby
safely reduce the numbers needing to be screened. In
Fig. 3, we model the period between screens assum-
ingaconstant breast cancer detectionrateof 4 1000.
The model predicts that high-risk women should be
screened every 15 months to achieve this rate of
detection whereas intermediate risk women would
need3 yearsandlowrisk6yearlyscreens.
Risk-adapted screening would require all women to
complete ariskassessment form. The PROCASstudy
outlined here indicates that only 40%of women who
attend for screening complete the form. Currently,
there is no incentive for women to join the study ex-
cept goodwill and the knowledge that they will be told
their risks. However, in the future, if women were
aware that the screening interval depended on risk,
this may be a greater incentive. Women not complet-
ing the risk formcould be maintained on the 3 years
interval.
Summary
The increasing incidence of breast cancer focuses on
the need for prevention and improved early detection
of the disease. FHCs are models for management of
younger women at risk but could potentially be used
for older women determined to be at high risk in the
UKNHSBSP.
Models such as Gail and TyrerCuzick predict gen-
eral risk well but have low discriminatory power for
individuals [73]. The models may (or may not) be
improved by adding other risk factors such as mam-
mographic density and measurement of breast
cancer riskassociated SNPs. Clinical trials of endo-
crine agents have demonstrated that it is possible to
prevent breast cancer using preventive therapies,
and observational studies suggest that lifestyle
changes may also reduce risk; however, ideally we
need appropriate randomized trials to test these
assumptions.
The early results of the Manchester PROCAS Study
suggests that it may be feasible to introduce risk pre-
diction and prevention strategies in the context of a
population-based mammographic screening pro-
gramme and thus to focus preventive approaches
andpossiblyintroducerisk-adaptedscreening.
Conict of internet statement
Noconict of interest wasdeclared.
Acknowledgements
We acknowledge the support of the Manchester Bio-
medical Research Centre and funding from the Na-
12
Current
screening
30%
20%
8
10
interval
6
L
i
f
e
t
i
m
e

r
i
s
k
B
r
e
a
s
t

c
a
n
c
e
r
s

p
e
r

1
0
0
0

p
e
r

y
e
a
r
10%
4
0
5%
2
0
Time from previous screen (years)
1 2 3 4 5 6
15 m* 70 m 36 m
Fig. 3 The rangeof mammographic density.
A. Howell et al.
|
Review: Prevention of breast cancer
2012 The Association for the Publication of the Journal of Internal Medicine 327
Journal of Internal Medicine, 2012, 271; 321330
tional Institute of Health Research (NIHR) and the
Genesis Breast Cancer Prevention Appeal. We would
like to thank the study radiologists and radiogra-
phers, the whole PROCAS team, Dr Stephen Eyre,
University of Manchester for advice with genotyping
andProfessor StephenDuffy for reviewing the manu-
script. This article presents, in part, independent re-
search commissioned by the National Institute for
Health Research (NIHR) under its Programme Grant
(Reference Number RP-PG-0707-10031). The views
expressedinthisarticleare thoseof the author(s) and
not necessarily those of the NHS, the NIHRor the UK
Department of Health.
References
1 CR-UK, CancerStats Incidence UK. Cancer ResearchUK2006.
CR-UK, CancerStatsIncidence UK. http://www.cancerresear-
chuk.org, 2009. http://info.cancerresearchuk.org/cancerstats/
types/breast/index.htm?script=trueaccessed02 01 2010.
2 Parkin DM, Bray F, Ferlay J, Pisani P. Global cancer statistics,
2002. CACancer JClin2005; 55: 74108.
3 Tryggvadottir L, SigvaldasonH, Olafsdottir GHet al. Population-
based study of changing breast cancer risk in Icelandic BRCA2
mutation carriers, 1920-2000. J Natl Cancer Inst 2006; 2:
11622.
4 Eaton SB, Pike MC, Short RV et al. Womens reproductive can-
cersinevolutionarycontext. QRevBiol 1994; 99: 35367.
5 Huang Z, Hankinson SE, Colditz GA et al. Dual effects of weight
and weight gain on breast cancer risk. JAMA 1997; 17:
140711.
6 Renehan AG, Soerjomataram I, Lietzman MF. Interpreting the
epidemiological evidence linking obesity and cancer: a frame-
work for population-attributable risk estimations in Europe.
Eu. J. Cancer 2010; 46: 258192.
7 Friedenreich CM. Physical activity and breast cancer: review of
theepidemiologicevidenceandbiologicmechanisms. Recent Re-
sultsCancer Res2011; 188: 12539.
8 Evans DG, Fentiman IS, McPherson K, Asbury D, Ponder BA,
Howell A. Familial breast cancer. BMJ1994; 6922: 1837.
9 Evans DGR, Cuzick J, Howell A. Cancer genetics clinics. Eur J
Cancer 1996; 32: 3912.
10 Parmigiani G, Berry DA, Aquilar O. Determining carrier proba-
bilities for breast cancer susceptibility genes BRCA1 and
BRCA2. AmJHumGenet 1998; 62: 1458.
11 Evans DG, Eccles DM, RahmanNet al. Anewscoring systemfor
the chances of identifying a BRCA1 2 mutation outperforms
existing models including BRCAPRO. J Med Genet 2004; 41:
47480.
12 Antoniou AC, Cunningham AP, Peto J et al. The BOADICEA
model of genetic susceptibility to breast and ovarian cancers:
updatesandextensions. Br JCancer 2008; 8: 145766.
13 Mavaddat N, RebbeckTR, Lakhani SR, EastonDF, AntoniouAC.
Incorporating tumour pathology information into breast cancer
riskpredictionalgorithms. Breast Cancer Res2010; 12: R28.
14 Evans DG, Lalloo F, Cramer Aet al. Additionof pathology andbi-
omarker information signicantly improves the performance of
the Manchester scoring system for BRCA1 and BRCA2 testing.
JMedGenet 2009; 46: 8117.
15 Gail MH, Brinton LA, Byar DP et al. Projecting individualized
probabilities of developing breast cancer for white females who
are being examined annually. J Natl Cancer Inst 1989; 81:
187986.
16 Tyrer J, Duffy SW, Cuzick J. A breast cancer prediction model
incorporating familial and personal risk factors. Stat Med 2004;
7: 111130.
17 Maurice A, Evans DG, Shenton A et al. Screening younger wo-
men with a family history of breast cancer does early detection
improveoutcome?Eur JCancer 2006; 42: 138590.
18 Domchek SM, Friebel TM, Singer CF et al. Association of risk
reducing surgery in BRCA1 or BRCA2 mutation carriers with
cancer riskandmortality. JAMA2010; 1: 96775.
19 FH01 Collaborative Teams. Mammographic surveillance in wo-
men younger than 50 years who have a family history of breast
cancer: tumour characteristics and projected effect onmortality
in the prospective, single-arm, FH01 study. Lancet Oncol 2010;
11: 112734.
20 Kurian AW, Sigal BM, Plevritis SK. Survival analysis of cancer
risk reduction strategies for BRCA1 2 mutation carriers. J Clin
Oncol 2010; 2: 22231.
21 Hagen AI, Kvistad KA, Maehle L et al. Sensitivity of MRI versus
conventional screening in the diagnosis of BRCA-associated
breast cancer in a national prospective series. Breast 2007; 16:
36774.
22 Costantino JP, Gail MH, Pee Det al. Validation studies for mod-
els projecting the risk of invasive and total breast cancer inci-
dence. JNatl Cancer Inst 1999; 91: 15418.
23 SpiegelmanD, Colditz GA, Hunter D, HertzmarkE. Validationof
the Gail et al. model for predicting individual breast cancer risk.
JNatl Cancer Inst 1994; 8: 6007.
24 Amir E, Evans DG, Shenton A et al. Evaluation of breast cancer
risk assessment packages in the family history evaluation and
screeningprogramme. JMedGenet 2003; 40: 80714.
25 Novotny J, Pecen L, Petruzelka L et al. Breast cancer risk
assessment in the Czech female population an adjustment of
the original Gail model. Breast Cancer Res Treat 2006; 95:
2935.
26 Decarli A, Calza S, Masala G, Specchia C, Palli D, Gail MH. Gail
model for prediction of absolute risk of invasive breast cancer:
independent evaluation in the Florence-European Prospective
Investigation Into Cancer and Nutrition cohort. J Natl Cancer
Inst 2006; 23: 168693.
27 Chlebowski RT, Anderson GL, Lane DS et al. Predicting risk of
breast cancer in postmenopausal women by hormone receptor
status. JNatl Cancer Inst 2007; 22: 1695705.
28 Amir E, Freedman OC, Seruga B, Evans DG. Assessing women
at highriskof breast cancer: areviewof riskassessment models.
JNatl Cancer Inst 2010; 102: 68091.
29 Jacobi CE, de Bock GH, Siegerink B, van Asperen CJ. Differ-
ences and similarities in breast cancer risk assessment models
in clinical practice: which model to choose? Breast Cancer Res
Treat 2009; 115: 38190.
30 McCormack VA, dos Santos Silva I. Breast density and paren-
chymal patterns as markers of breast cancer risk: a meta-analy-
sis. Cancer Epidemiol BiomarkersPrev2006; 15: 115969.
31 CummingsSR, TiceJA, Bauer Set al. Preventionof breast cancer
in postmenopausal women: approaches to estimating and
reducingrisk. JNatl Cancer Inst 2009; 6: 38498.
32 BarlowWE, White E, Ballard-BarbashRet al. Prospective breast
cancer risk prediction for women undergoing screening mam-
mography. JNatl Cancer Inst 2006; 17: 120414.
A. Howell et al.
|
Review: Prevention of breast cancer
328 2012 The Association for the Publication of the Journal of Internal Medicine
Journal of Internal Medicine, 2012, 271; 321330
33 Tice JA, Cummings SR, Smith-Bindman R, Ichikawa L, Barlow
WE, Kerlikowske K. Using clinical factors and mammographic
breast density to estimate breast cancer risk: development and
validation of a new predictive model. Ann Intern Med 2008; 5:
33747.
34 Chen J, Pee D, Ayyagari R et al. Projecting absolute invasive
breast cancer risk in white women with a model that includes
mammographicdensity. JNCI 2006; 98: 121526.
35 Stacey SN, Manolescu A, Sulem P et al. Common variants
on chromosomes 2q35 and 16q12 confer susceptibility to
estrogen receptor-positive breast cancer. Nat Genet 2007; 39:
8659.
36 Hunter DJ, Kraft P, Jacobs KB et al. Agenome-wide associ-
ation study identies alleles in FGFR2 associated with risk of
sporadic postmenopausal breast cancer. Nat Genet 2007; 39:
8704.
37 Zheng W, Long J, Gao YT et al. Genome-wide association study
identies a new breast cancer susceptibility locus at 6q25.1.
Nat Genet 2009; 41: 3248.
38 Turnbull C, AhmedS, MorrisonJet al. Genome-wideassociation
study identies ve new breast cancer susceptibility loci.
Nat Genet 2010; 42: 5047.
39 vanZitterenM, vander Net JB, KunduS, FreedmanAN, vanDui-
jn CM, Janssens AC. Genome-based predictionof breast cancer
risk inthe general population: a modeling study based on meta-
analyses of genetic associations. Cancer Epidemiol Biomarkers
Prev2011; 20: 922.
40 Gail MH. Discriminatory accuracy from single-nucleotide poly-
morphismsinmodelstopredict breast cancer risk. JNatl Cancer
Inst 2008; 14: 103741.
41 Gail MH. Value of adding single-nucleotide polymorphismgeno-
types to a breast cancer risk model. J Natl Cancer Inst 2009; 13:
95963.
42 Comen E, Balistreri L, Gonen M et al. Discriminatory accuracy
and potential clinical utility of genomic proling for breast can-
cer risk in BRCA-negative women. Breast Cancer Res Treat
2011; 127: 47987.
43 Mealiffe ME, Stokowski RP, Rhees BK, Prentice RL, Pettinger M,
Hinds DA. Assessment of clinical validity of a breast cancer risk
model combining genetic and clinical information. J Natl Cancer
Inst 2010; 102: 161827.
44 ParkJH, Wacholder S, Gail MHet al. Estimationof effect sizedis-
tribution from genome-wide association studies and implica-
tionsfor futurediscoveries. Nat Genet 2010; 42: 5705.
45 Chatterjee N, Park JH, Caporaso N, Gail MH. Predicting the
future of genetic risk prediction. Cancer Epidemiol Biomarkers
Prev2011; 20: 38.
46 Pharoah PD, Antoniou AC, Easton DF, Ponder BA. Polygenes,
risk prediction, and targeted preventionof breast cancer. NEngl
JMed2008; 26: 2796803.
47 Milne RL, Gaudet MM, Spurdle AB et al. Assessing interactions
between the associations of common genetic susceptibility
variants, reproductive history and body mass index with breast
cancer risk in the breast cancer association consortium: a
combined case-control study. Breast Cancer Res 2010; 12:
R110.
48 Travis RC, Reeves GK, Green J et al. Gene-environment
interactions in 7610 women with breast cancer: prospective
evidence from the Million Women Study. Lancet 2010; 9732:
214351.
49 Santen R, Boyd N, Chlebowski RT et al. Critical assessment of
new risk factors for breast cancer: considerations for develop-
ment of animprovedrisk predictionmodel. Endocr Relat Cancer
2007; 14: 16987.
50 Cuzick J, DeCensi A, Arun Bet al. Preventive therapy for breast
cancer: a consensus statement. Lancet Oncol 2011; 12: 496
503.
51 Cuzick J, Powles T, Veronesi U et al. Overview of the main out-
comes in breast-cancer prevention trials. Lancet 2003; 9354:
296300.
52 Vogel VG, Costantino JP, WickerhamDL et al. Update of the na-
tional surgical adjuvant breast and bowel project Study of
Tamoxifen and Raloxifene (STAR) P-2 trial: preventing breast
cancer. Cancer PrevRes(Phila) 2010; 3: 696706.
53 Goss PE, Ingle JN, Ales-Mart nez JE et al. Exemestane for
breast-cancer prevention in postmenopausal women. N Engl J
Med2011; 25: 238191.
54 Cuzick J. IBIS II: a breast cancer prevention trial in postmeno-
pausal women using the aromatase inhibitor anastrozole. Ex-
pert RevAnticancer Ther 2008; 8: 137785.
55 Fisher B, Costantino JP, WickerhamDL et al. Tamoxifen for the
preventionof breast cancer: current status of the national surgi-
cal adjuvant breast and bowel project P-1 study. J Natl Cancer
Inst 2005; 22: 165262.
56 Martino S, Cauley JA, Barrett-Connor E et al. Continuing out-
comes relevant to Evista: breast cancer incidence in postmeno-
pausal osteoporotic women in a randomized trial of raloxifene.
JNatl Cancer Inst 2004; 23: 175161.
57 Barrett-Connor E, Mosca L, Collins P et al. Effects of raloxifene
on cardiovascular events and breast cancer in postmenopausal
women. NEngl JMed2006; 2: 12537.
58 Cuzick J, Forbes JF, Sestak I et al. Long-termresults of tamoxi-
fen prophylaxis for breast cancer 96-month follow-up of the
randomizedIBIS-I trial. JNatl Cancer Inst 2007; 4: 27282.
59 Freedman AN, Yu B, Gail MH et al. Benet Riskassessment
for breast cancer chemoprevention with raloxifene or tamoxifen
for women age 50 years or older. J Clin Oncol 2011; 17:
232733.
60 Cuzick J, Warwick J, Pinney E et al. Tamoxifen-induced
reduction in mammographic density and breast cancer risk
reduction: a nested case-control study. J Natl Cancer Inst 2011;
9: 74452.
61 Harvie M, Howell A, Vierkant RA et al. Association of gain and
lossof weight beforeandafter menopausewithriskof postmeno-
pausal breast cancer in the Iowa womens health study. Cancer
Epidemiol BiomarkersPrev2005; 14: 65661.
62 Eliassen AH, Colditz GA, Rosner B, Willett WC, Hankinson SE.
Adult weight change and risk of postmenopausal breast cancer.
JAMA2006; 2: 193201.
63 Byers T, Sedjo RL. Does intentional weight loss reduce cancer
risk?DiabetesObesMetab2011; 13: 106372.
64 Teras LR, GoodmanM, Patel AV, Diver WR, FlandersWD, Feigel-
sonHS. Weight loss andpostmenopausal breast cancer inapro-
spective cohort of overweight and obese US women. Cancer
CausesControl 2011; 22: 5739. Epub2011Feb13.
65 Prentice RL, Caan B, Chlebowski RT et al. Low-fat dietary pat-
tern and risk of invasive breast cancer: the womens health ini-
tiative randomized controlled dietary modication trial. JAMA
2006; 6: 62942.
66 Bennett RL, Sellars SJ, Moss SM. Interval cancers in the NHS
breast cancer screening programme in England, Wales and
NorthernIreland. Br JCancer 2011; 4: 5717.
67 KirshVA, Chiarelli AM, Edwards SAet al. Tumor characteristics
associatedwithmammographicdetectionof breast cancer inthe
A. Howell et al.
|
Review: Prevention of breast cancer
2012 The Association for the Publication of the Journal of Internal Medicine 329
Journal of Internal Medicine, 2012, 271; 321330
ontario breast screening program. J Natl Cancer Inst 2011; 12:
94250.
68 Gilbert FJ, Astley SM, McGee MAet al. Single reading with com-
puter aided detection and double reading of screening mammo-
grams in the United Kingdom National Breast Screening
Program. Radiology2006; 241: 4753.
69 BoydNF, MartinLJ, Yaffe M, MinkinS. Mammographic density.
Breast Cancer Res2009; 11(Suppl. 3): S4.
70 Shepherd JA, Kerlikowske K, Ma L et al. Volume of mammo-
graphic density and risk of breast cancer. Cancer Epidemiol Bio-
markersPrev2011; 20: 147382.
71 MandelsonMT, Oestreicher N, Porter PLet al. Breast densityasa
predictor of mammographic detection: comparison of interval-
and screen-detected cancers. J Natl Cancer Inst 2000; 92:
10817.
72 NixonRM, PharoahP, Tabar Let al. Mammographicscreeningin
women witha family history of breast cancer: some results from
the Swedish two-county trial. Rev Epidem et Sante Publ 2000;
48: 32531.
73 Rockhill B, SpiegelmanD, Byrne C, Hunter DJ, Colditz GA. Vali-
dationof theGail et al. model of breast cancer riskpredictionand
implications for chemoprevention. J Nat Cancer Inst 2001; 93:
35866.
Correspondence: A. Howell, Genesis Prevention Centre and Nightin-
gale Breast ScreeningCentre, UniversityHospital of SouthManches-
ter, Southmoor Road, Wythenshawe, Manchester, M239LT, UK.
(e-mail: anthony.howell@christie.nhs.uk).
A. Howell et al.
|
Review: Prevention of breast cancer
330 2012 The Association for the Publication of the Journal of Internal Medicine
Journal of Internal Medicine, 2012, 271; 321330

Potrebbero piacerti anche