You are on page 1of 16

REVIEWS

Epidemiology of Alzheimer disease


Christiane Reitz, Carol Brayne and Richard Mayeux
Abstract | The global prevalence of dementia is estimated to be as high as 24 million, and is predicted to
double every 20 years through to 2040, leading to a costly burden of disease. Alzheimer disease (AD) is
the leading cause of dementia and is characterized by a progressive decline in cognitive function, which
typically begins with deterioration in memory. Before death, individuals with this disorder have usually become
dependent on caregivers. The neuropathological hallmarks of the AD brain are diffuse and neuritic extracellular
amyloid plaques—which are frequently surrounded by dystrophic neurites—and intracellular neurofibrillary
tangles. These hallmark pathologies are often accompanied by the presence of reactive microgliosis and the
loss of neurons, white matter and synapses. The etiological mechanisms underlying the neuropathological
changes in AD remain unclear, but are probably affected by both environmental and genetic factors. Here, we
provide an overview of the criteria used in the diagnosis of AD, highlighting how this disease is related to, but
distinct from, normal aging. We also summarize current information relating to AD prevalence, incidence and
risk factors, and review the biomarkers that may be used for risk assessment and in diagnosis.
Reitz, C. et al. Nat. Rev. Neurol. 7, 137–152 (2011); published online 8 February 2011; doi:10.1038/nrneurol.2011.2

most frequent cause of dementia—is associated with an


Continuing Medical Education online
estimated health-care cost of US$172 billion per year.1
This activity has been planned and implemented in accordance The key pathological changes that are observed in AD
with the Essential Areas and policies of the Accreditation Council brain tissue are increased levels of both the amyloid‑β
for Continuing Medical Education through the joint sponsorship of
Medscape, LLC and Nature Publishing Group. Medscape, LLC is
(Aβ) peptide, which is deposited extracellularly in
accredited by the ACCME to provide continuing medical education diffuse and neuritic plaques, and hyperphosphory-
for physicians. lated tau (p-tau), a microtubule assembly protein that
Medscape, LLC designates this Journal-based CME for a maximum accumulates intra­cellularly as neurofibrillary tangles
of 1 AMA PRA Category 1 CreditsTM. Physicians should claim only (NFTs). In addition to these pathologies, widespread loss
the credit commensurate with their participation in the activity.
of neurons and synapses is observed. The mechanisms
All other clinicians completing this activity will be issued a
certificate of participation. To participate in this journal CME
underlying the changes outlined above are unknown but,
activity: (1) review the learning objectives and author disclosures; when present, these neuro­pathological features confer a
(2) study the education content; (3) take the post-test and/or diagnosis of definite AD, according to current diagnostic
complete the evaluation at http://www.medscapecme.com/ criteria (see below).
journal/nrneuro; (4) view/print certificate.
In this article, we review the criteria used to diagnose
Released: 8 February 2011; Expires: 8 February 2012
AD, and summarize current knowledge regarding AD
Learning objectives prevalence, incidence, and genetic and nongenetic risk
Upon completion of this activity, participants should be able to: factors. We also examine the value and limitations of bio-
1 Describe current information regarding AD prevalence, markers that may be used to determine AD risk and to
incidence, and risk factors.
2 Describe criteria used in the diagnosis of AD and how it is
aid diagnosis of this disease.
related to but distinct from normal aging. Gertrude H. Sergievsky
3 Describe biomarkers that may be used for risk assessment Diagnostic criteria Center, Columbia
and in the diagnosis of AD. University Medical
Over the past century, the classification of dementia sub- Center, 630 West 168th
types has been revised repeatedly (Box 1). The key classi­ Street, New York,
Introduction fication for the diagnosis of AD has been the National NY 10032, USA
(C. Reitz, R. Mayeux).
Dementia is characterized by deterioration in cognition, Institute of Neurological and Communicative Disorders Department of Public
function and behavior, and places a considerable burden and Stroke and Alzheimer’s Disease and Related Health and Primary
Care, Institute of
on society. In the US alone, Alzheimer disease (AD)—the Disorders Association (NINCDS–ADRDA; Box 2) cri- Public Health,
teria, which were established in 1984.1 These criteria link University Forvie Site,
clinical patterns to neuropathology and use the labels Robinson Way,
Competing interests
Cambridge CB2 0SR,
R. Mayeux declares an association with the following possible, probable and definite AD in patient diagnosis.2 UK (C. Brayne).
organization: NIH. See the article online for full details of the
Despite the widespread use of this classification, several
relationship. The other authors, the journal Chief Editor Correspondence to:
H. Wood and the CME questions author L. Barclay declare no notable issues are associated with the extrapolation of R. Mayeux
competing interests. clinicopathological findings for diagnosis (Box 3). rpm2@columbia.edu

NATURE REVIEWS | NEUROLOGY VOLUME 7  |  MARCH 2011  |  137


© 2011 Macmillan Publishers Limited. All rights reserved
REVIEWS

Key points Since its instigation, the definition of an intermediate


■■ The unprecedented level of aging occurring in developed nations will lead to an
state between normal cognition and dementia, namely
enormous burden of Alzheimer disease (AD) mild cognitive impairment (MCI), has been modified
■■ The primary pathological hallmarks in AD brain tissue—diffuse and neuritic
and investigated in many settings. In clinical settings,
extracellular amyloid plaques and intracellular neurofibrillary tangles—are well where cognitive impairment is most likely to be detected,
known, but the underlying etiologies of these pathologies remain unclear MCI has proved a useful label to define people who are at
■■ The diagnosis of AD in living patients is based on clinical examination—no risk of developing AD. In population settings, however,
definite diagnostic test is currently available—but may be supported by the use where cognitive impairment is less likely to be detected,
of clinical biomarkers this label has proved to be less valuable.
■■ AD heritability varies from 58–79% depending on age at onset; however, only
a portion of the likely substantial genetic contribution to this disease has Prevalence and incidence
been determined By 2005, 24.2 million people worldwide had dementia
■■ Several nongenetic factors (including recognized vascular risk factors) have and 4.6 million new cases of this condition were arising
been associated with AD, but the underlying mechanisms linked to these every year (Table 1);3 ≈ 70% of these cases were attributed
factors are uncertain to AD. Among regional populations of individuals aged
≥60 years, those from North America and Western Europe
exhibited the highest prevalence of dementia (6.4% and
Box 1 | Historical perspective on the AD diagnosis 5.4%, respectively), followed by those from Latin America
(4.9%) and China and its western-Pacific neighbours
1906–1950
During this period, individuals aged <65 years who developed dementia were
(4.0%). Meanwhile, the annual regional dementia inci-
diagnosed as having AD or presenile dementia when no other known causes of dence rates (per 1,000 individuals in the population) were
dementia were present. Individuals with dementia who had a strong history of estimated to be 10.5 for North America, 8.8 for Western
vascular disease and exhibited executive symptoms were labeled as having multi- Europe, 9.2 for Latin America, and 8.0 for China and its
infarct dementia, independent of their age. Patients who developed dementia western-Pacific neighbours.3 For all these populations,
aged >65 years and for whom no other cause of this condition was known were the incidence rate for dementia increased exponentially
diagnosed as having senile dementia. This form of dementia was understood to with age, with the most notable rise occurring through
be vascular and prompted the use of vasodilators.
the seventh and eighth decades of life. The prevalence and
1950–1980 incidence rates for AD also increase exponentially with
In this period, late-onset AD was recognized to be pathologically indistinguishable age (Supplementary Figure 1 online).
from the early-onset form of this disease. In addition, the vascular explanation for The incidence rates of AD and dementia in people
senile dementia was largely abandoned. Consequently, the age criterion for the
aged <75 years seem to be relatively similar across
label AD was lifted, the term senile dementia was abandoned, and the diagnosis
of AD was assigned when all other possible causes of cognitive impairment were
studies, but in the oldest age groups these rates vary
excluded. In addition, a key publication by Katman and colleagues noted the (Supplementary Figure 1 online). Methodological issues
“malignancy” of AD.205 partly account for the observed divergence; however, the
estimates might also reflect geographical differences in
1980–present
age-dependent incidence owing to variation in survival
The vascular component of late-onset AD has again been recognized. In 1984,
the National Institute of Neurological and Communicative Disorders and Stroke and the prevalence of risk and protective factors.
and Alzheimer’s Disease and Related Disorders Association criteria were
implemented. The key to these criteria is the linkage of clinical syndromal Risk and protective factors
patterns to neuropathology after death through the use of possible, probable and Various risk and protective factors have been linked to
definite labels for AD.2 dementia and/or AD. Two types of analytical approaches
Abbreviation: AD, Alzheimer disease.
have been used to identify such factors, namely obser-
vational studies (that is, case–control, cohort and cross-
­s ectional studies) and experimental studies (that is,
The original NINCDS–ADRDA criteria have received clinical trials). The strengths and weaknesses of such
updates and undergone revisions, and have been incor- approaches are highlighted in Table 2.
porated into major international criteria, including the
Diagnostic and Statistical Manual of Mental Disorders Risk factors
(DSM; Box 2) and the International Classification of Risk factors are antecedents of part of the disease pathway,
Diseases. The NINCDS–ADRDA and DSM criteria and can be associated with the etiology or the outcome of
are again under review, as the AD spectrum is now a disease. Such factors may be used to assess disease risk
understood to be broader than was previously thought: but do not usually provide sufficient sensitivity and/or
pathological changes (such as cerebrovascular changes) specificity to be employed as diagnostic markers.
other than the key AD pathologies (that is, amyloid Various risk factors have been found to be associated
plaques and NFTs) are now believed to precede or with dementia and/or AD. Of note, many recognized vas-
coexist with AD, and contribute to the cognitive and cular risk factors for ischemic heart disease and/or stroke
physical dysfunction observed in this disorder. The are also risk factors for dementia. Diabetes, hyper­tension,
use of biomarkers may increase diagnostic specificity smoking and obesity have all been found to increase
for AD and, consequently, will be considered in the dementia risk. Nevertheless, while vascular risk factors and
updated criteria. cerebrovascular disease clearly underlie vascular dementia,

138  |  MARCH 2011  |  VOLUME 7  www.nature.com/nrneurol


© 2011 Macmillan Publishers Limited. All rights reserved
REVIEWS

an etiological role for vascular changes in Aβ deposition Box 2 | The NINCDS–ADRDA criteria
and, hence, AD remains unclear. Figure 1 shows possible
The NINCDS–ADRDA criteria categorize AD as definite, probable or possible.2 A
mechanisms linking vascular disease to dementia.
diagnosis of definite AD requires supporting pathological evidence. Probable AD is
the maximum level of certainty possible without pathological confirmation. This
Cerebrovascular disease category requires a gradual onset and progressive decline in memory with
Cerebrovascular changes such as hemorrhagic infarcts, involvement of at least one other cognitive domain that is established by clinical
small and large ischemic cortical infarcts, vasculo­ examination and confirmed by neuropsychological tests. The NINCDS–ADRDA
pathies, and white matter changes all increase the risk criteria further require that the patient is fully conscious and does not have another
of dementia. A meta-analysis incorporating data from condition that might explain their symptoms. Other supportive features of probable
22 hospital-­based and eight population-based cohorts4 AD include evidence of progressive deterioration of language, praxis and visual
recognition, with impaired activities of daily living or a positive family history of AD.
found that 7.4% of patients with first-ever stroke devel-
Normal levels of routine cerebrospinal fluid measures, evidence of progressive
oped poststroke dementia. Several mechanisms exist cerebral atrophy on brain imaging and a normal or nonspecific pattern on EEG
through which stroke could lead to cognitive impair- support probable AD diagnosis. Features that make the diagnosis of probable AD
ment and AD. First, stroke may lead directly to damage unlikely include an apoplectic onset, the presence of focal neurological findings
of brain regions that are important in memory function, (such as hemiparesis or sensory loss), and the presence of seizures or gait
such as the thalamus and the thalamocortical projec- disturbances. Patients meet criteria for possible AD when the course of cognitive
tions. Second, stroke might increase Aβ deposition, decline is atypical, focal neurological findings are evident, or disorders coexist that
by themselves can explain the dementia, such as stroke or traumatic brain injury.
which in turn can lead to cognitive decline. Third, the
The DSM-IV criteria for AD require a gradual onset and progressive impairment
onset of stroke may induce inflammatory responses that in memory function and at least one other cognitive domain that results in
impair cognitive function. Last, hypoperfusion can lead impairment of social and occupational function. Such cognitive impairment would
to overexpression of cyclin-dependent kinase 5 (CDK5), not be explained by other psychiatric, neurological or systemic diseases. The
a serine–threonine kinase that is critical to synapse for- NINCDS–ADRDA and DSM criteria are currently being revised.
mation and synaptic plasticity and, hence, to learning
Abbreviations: AD, Alzheimer disease; DSM, Diagnostic and Statistical Manual of Mental
and memory.5 Aberrant CDK5 activation is associated Disorders; NINCDS–ADRDA, National Institute of Neurological and Communicative Disorders
with neuronal apoptosis and death. 6 This kinase may and Stroke and the Alzheimer’s Disease and Related Disorders Association.
also be involved in the abnormal phosphorylation of
tau, thereby contributing to the formation of NFTs,7
and might be a key protein linking NFT pathology to Box 3 | Issues relating to the diagnosis of AD
amyloid plaques.
Two notable issues surround the use of neuropathological
studies to inform the diagnosis of AD. First,
Blood pressure
neuropathological studies are usually end-stage disease
Inconsistencies exist between data from cross-sectional investigations and extrapolation of the findings of such
and longitudinal studies that have examined the effect of studies to living patients in earlier stages of disease
blood pressure on brain function. In part, these inconsis- can be difficult. Second, neuropathological findings
tencies can be attributed to differences in study design; can differ between patients with similar clinical profiles
specifically, variation in the time between measurement before death. Thus, clinical diagnoses are not always
of blood pressure and assessment of cognitive abilities, in agreement with clincopathological diagnoses, which
and in the age at which these parameters were measured. indicates that the relationship between clinical profiles
and age is not stable. Estimates of AD prevalence and
In contrast to the results of cross-sectional and longitu-
incidence will continue to vary until the key underlying
dinal studies, data from observational studies exploring pathologies of the disease have been identified, the
the association between elevated levels of blood pressure relationship of specific pathology to clinical phenotype is
in midlife (40–60 years of age) and late-life cognitive understood, and the clinical diagnosis of AD is improved.
impairment have proved to be relatively consistent across
Abbreviation: AD, Alzheimer disease.
cohorts,8–11 although results relating to the associ­ation
between late-life blood pressure levels and cognitive
decline and dementia remain inconsistent.12–15 result of vessel stiffening, weight loss and changes in the
In middle age, elevated blood pressure increases autonomic regulation of blood flow.
the risk of cognitive impairment, dementia and AD. Seven randomized, placebo-controlled trials (RCTs)
Hypertension may increase the risk of AD by decreasing have evaluated the benefit of antihypertensive treat-
the vascular integrity of the blood–brain barrier (BBB), ments in patients with cognitive impairment. Three
resulting in protein extravasation into brain tissue.16 studies showed that such agents had a beneficial effect.
In turn, protein extravasation can lead to cell damage, In the Syst-Eur trial, which included patients without
a reduction in neuronal or synaptic function, apop­ dementia who were aged >60 years,18 the risk of demen-
tosis, and an increase in Aβ accumulation, resulting in tia (diagnosed according to the revised DSM-III criteria)
cognitive impairment.17 With increasing age, the effect was reduced by 55% in the treatment group (nitren­
of elevated blood pressure on AD risk diminishes and dipine alone or in combination with enalapril maleate,
may even become inverted, with an increase in blood hydrochlorothiazide or both add-on drugs) compared
pressure showing a protective effect. This observation with the placebo group. In PROGRESS,19 which involved
might be explained by the fact that following the onset 6,105 individuals with prior stroke or transient ischemic
of AD, blood pressure begins to decrease, possibly as a attack, patients treated with perindopril (with or without

NATURE REVIEWS | NEUROLOGY VOLUME 7  |  MARCH 2011  |  139


© 2011 Macmillan Publishers Limited. All rights reserved
REVIEWS

Table 1 | Prevalence and incidence of dementia in developed and developing regions


Region Consensus Estimated annual People with Estimated increase in
dementia incidence of dementia aged proportion of people
prevalence at age dementia (per ≥60 years in with dementia from
≥60 years (%) 1,000 individuals) 2001 (millions) 2001 to 2040 (%)
Western Europe 5.4 8.8 4.9 102
Eastern Europe (regions with low adult 3.8 7.7 1.0 169
mortality)
Eastern Europe (regions with high adult 3.9 8.1 1.8 84
mortality)
North America 6.4 10.5 3.4 172
Latin America 4.6 9.2 1.8 393
North Africa and Middle Eastern 3.6 7.6 1.0 385
Crescent
Developed western Pacific 4.3 7.0 1.5 189
China and developing western Pacific 4.0 8.0 6.0 336
Indonesia, Thailand and Sri Lanka 2.7 5.9 0.6 325
India and south Asia 1.9 4.3 1.8 314
Africa 1.6 3.5 0.5 235
Combined values 3.9 7.5 24.3 234
Data taken from Ferri et al. (2005).3

indapamide) saw a 12% reduction in their risk of demen- Diabetes and impairment of glucose tolerance lead to
tia (diagnosed according to DSM-IV criteria) in the the formation of advanced glycosylation end products
absence of recurrent stroke, a 34% decrease in their risk (AGEs), and amyloid plaques and NFTs contain recep-
of dementia with recurrent stroke, and a 54% drop in tors for AGEs (RAGEs). Glycation of Aβ enhances its
their risk of cognitive decline (decline of ≥3 points in the propensity to aggregate in vitro.30 In addition, RAGEs
Mini Mental State Examination [MMSE]) with recurrent may facilitate the neuronal damage caused by Aβ,31 as the
stroke, in comparison with patients who received latter is a high-affinity ligand for cell-surface RAGEs.
placebo. Among 81 community-screened, dementia- Adipose tissue produces adipokines (such as adipo­
free individuals (aged >69 years, with no previous record nectin and leptin) and cytokines (including resistin,
of receiving antihypertensive treatment) who under­ tumor necrosis factor [TNF] and interleukin [IL]‑6),
went captopril or bendrofluazid treatment in the which are involved in metabolism and inflammation,
HOPE trial,20 those in the most treatment-responsive respectively. Adipokine and cytokine levels increase with
quartile (blood pressure reduction ≥19 mmHg) had insulin resistance and hyperinsulinemia. Nevertheless,
improved scores on the Anomalous Sentences and whether adipokines and cytokines are indicators of insulin
Paired Associates Tests compared with those in the least resistance remains unclear, as is whether these signaling
treatment-responsive quartile (blood pressure reduction molecules are causally related to AD. Evidence suggests
≤5 mmHg). Nevertheless, other RCTs evaluating the that leptin exerts beneficial effects on brain regions that
benefit of antihypertensive treatment on cognitive are important for memory function, including the CA1
decline (that is, the MRC study, 21 SHEP, 22 SCOPE 23 region of the hippocampus.32 Moreover, leptin receptor-
and HYVET-COG 24) have disclosed nonsignificant deficient transgenic mice showed less synaptic plasticity
effects from such therapies, thereby failing to clarify and showed poorer performance on spatial memory tasks
the effect of blood-pressure-lowering medication on than did wild-type mice33 and, in the Framingham study,
cognitive decline. high circulating plasma leptin levels were associated with
a decrease in AD risk and high cerebral brain volumes
Type 2 diabetes in cognitively normal adults, although these associations
In observational studies, type 2 diabetes (T2D) has were restricted to non-obese people.34
been found to nearly double the risk of AD.25–27 Various A meta-analysis by Profenno et al., which included 10
mechanisms have been proposed whereby diabetes might longitudinal studies examining the relationship between
influence the development of AD. In cases of hyper­ T2D and AD,35 found that T2D increased the risk of AD
insulinemia accompanying diabetes, insulin may compete by 54%. Nevertheless, studies examining the relation-
with Aβ for the insulin degrading enzyme (IDE), thereby ship between T2D and the neuropathological hallmarks
hindering clearance of Aβ from the brain.28 Moreover, of AD have proved inconsistent: while some studies
a histopathological study of hippocampal tissue from showed that patients with diabetes had higher numbers
patients with AD and healthy controls showed relative of amyloid plaques or NFTs than healthy controls,36,37
reductions in IDE expression and IDE messenger RNA other investigations found no association or an inverse
levels in AD brain tissue.29 relationship between such lesions and diabetes.37,38

140  |  MARCH 2011  |  VOLUME 7  www.nature.com/nrneurol


© 2011 Macmillan Publishers Limited. All rights reserved
REVIEWS

Table 2 | Common types of epidemiological study design


Study type Methodological approach Advantages Disadvantages
Nonexperimental studies
Case–control study Sampling conducted with respect Low costs in relation to efficiency Disease and exposure status determined
to disease status simultaneously, making the temporal sequence of
Estimates the odds of having been events often difficult to establish
exposed to a risk factor given the
current case–control status
Cross-sectional study Includes all individuals of a Low costs in relation to efficiency Disease and exposure status determined
population regardless of exposure simultaneously, making the temporal sequence of
or disease status events often difficult to establish
Overrepresentation of cases with long duration and
under-representation of cases with short duration of
illness, leading to bias
Cohort study Sampling conducted with respect Exposure status is ascertained before Requires the follow-up of a large number of individuals
to exposure status the occurrence of disease, allowing until disease development and, hence, is costly
incidence rates of disease to be Difficult to control confounding variables and
calculated in people with and without maintain high follow-up rates
risk factors Adverse outcomes may occur before the onset of the
disease of interest, leading to survival bias
Experimental (randomized) studies
Clinical trial Individuals are randomly assigned to Minimizes bias and allows valid Participants studied may not be representative of
intervention and comparison groups statistical testing, as factors that could the general population
Aims to evaluate a cure or a potentially confound the examined
preventive treatment (usually drugs) association should occur with roughly
equal frequencies in the intervention
and comparison groups

Several RCTs have explored the effects of anti­diabetic inflammatory gene expression, alter Aβ homeostasis
drugs on cognitive impairment. In a study by Reger and exhibit neuro­protective effects.43
and colleagues,39 intranasal administration of insulin
altered cognitive performance in elderly patients with Body weight
amnestic MCI or early AD in an apolipoprotein E Early studies examining body fat distribution and
(APOE) genotype-­d ependent fashion. In memory- cognitive dysfunction showed that low BMI or being
impaired adults who did not harbor the APOE ε4 allele underweight seemed to be risk factors for dementia
(a genetic risk factor for AD; see below), but not in APOE and age-related brain changes such as atrophy.44,45 Later
ε4-positive adults, insulin administration facilitated pros­pective studies, however, have linked both low and
immediate recall at doses of 20 IU (P = 0.0006) and 40 IU high body weight to a heightened risk of AD and cog-
(P = 0.0013) compared with placebo treatment. In a study nitive impairment, suggesting a U‑shaped relationship
of patients with AD or amnestic MCI by Watson and col- between weight and cognitive performance.46–48 The
leagues, which considered effect sizes (f 2) of 0.02, 0.15 association of body weight with the risk of AD seems
and 0.35 as small, medium and large effects, respectively, to depend on the age at which body weight is measured.
individuals who received 6 months of treatment with the In addition, evidence exists for reverse causation in the
peroxisome proliferator-activated receptor (PPAR)‑γ years preceding dementia onset; that is, loss of body
agonist rosiglitazone (n = 20) showed moderate-­to-large weight is caused by cognitive impairment during the
improvements in delayed recall (P = 0.0138; f 2 = 0.31) and prodromal phase of dementia.49 Of note, the relation-
large improvements in selective attention (P = 0.0061; ship between body weight and cognitive impairment
f 2 = 0.38) compared with patients receving placebo and dementia seems to be driven by central obesity.50
(n = 10). 40 A study of 511 patients with AD demon- The aforementioned meta-analysis by Profenno et al.
strated that APOE ε4 noncarriers exhibited cognitive showed that obsesity (as assessed by BMI) increased
and functional improvement in response to rosiglitazone the risk of AD by 59%.35
(mean changes in the AD Assessment Scale cognition
subscale [ADAS-cog] for placebo and 8 mg rosiglita- Plasma lipid levels
zone were 1.5 and –2.0, respectively; P = 0.02).41 Finally, Conflicting data are available concerning the relation-
Sato et al. showed that daily treatment with 15–30 mg ship between dyslipidemia and cognitive impairment or
pioglitazone was associated with improvements in AD.51–56 Amyloid precursor protein (APP) can be broken
scores on the MMSE (23.1 ± 4.1 with pioglitazone down by enzymes, termed the secretases, via two routes,
versus 22.1 ± 3.5 with placebo) and on the ADAS-cog the nonamyloidogenic and amyloidogenic pathways. In
(142 ± 6.5 with pioglitazone versus 15.5 ± 5.9 with the second pathway, APP is proteolytically cleaved by
placebo) after 6 months (n = 21 in both treatment arms).42 β‑secretase and, subsequently, γ‑secretase to gene­rate
Of note, PPARγ agonists have been shown to inhibit Aβ, the most common isoforms of which comprise

NATURE REVIEWS | NEUROLOGY VOLUME 7  |  MARCH 2011  |  141


© 2011 Macmillan Publishers Limited. All rights reserved
REVIEWS

Vascular cognitive
Cerebrovascular disease impairment
(infarcts, white matter lesions) Memory impairment
(retrieval)
Dysexecutive syndrome
Diabetes Vascular syndrome
Hypertension
Obesity
Dyslipidemia
Metabolic syndrome Alzheimer disease
Blood–brain barrier dysfunction
Oxidation Memory impairment
Hyperinsulinemia Amyloid-β deposition in brain (delayed recall)
Adipokines Amnestic mild cognitive
Cytokines impairment
Dementia

Figure 1 | Potential mechanisms linking vascular risk factors and cognitive impairment. At least two pathways exist that
result in cognitive impairment and dementia: development of cerebrovascular disease may lead to vascular cognitive
impairment syndromes, and deposition of amyloid‑β may lead to other distinct amnestic clinical syndromes, including
Alzheimer disease. In addition, these pathways may overlap and interact, resulting in mixed cognitive syndromes.

40 (Aβ1–40) and 42 (Aβ1–42) amino acids, with the latter the inflammatory immune system, leading to activation
being the most fibrogenic of the two peptide species. of phagocytes and further oxidative damage.74 In addi-
Evidence exists that depletion of membrane cholesterol tion, smoking may promote cerebrovascular disease.
inhi­bits secretase cleavage of APP, thereby lowering Evidence also exists, however, that smoking can have a
Aβ 1–40 and Aβ 1–42 accumulation. Nevertheless, dys­ protective effect against AD. Nicotine has been sug-
lipidemia increases the risk of vascular disease, which gested to induce an increase in the level of nicotinic
in turn is associated with a heightened risk of AD. In acetyl­c holine receptors, thereby counterbalancing
people at risk of cardiovascular and cerebrovascular the loss of these receptors, and subsequent cholinergic
disease, statins are the first-line treatments for reduc- deficits, observed in AD.75
ing cholesterol levels. Three RCTs—involving 748 par-
ticipants aged 50–90 years—have explored the effect of Depressive symptoms
statins on AD risk.58–60 Overall, these studies yielded Depressive symptoms occur in 40–50% of patients with
insufficient evidence for a beneficial effect of statins on AD. Some longitudinal and case–control studies have
AD risk. The results of a large-scale trial of simvastatin found an increase in the risk of AD or MCI in individu-
to slow AD progression have yet to be published.61 als with a history of depression,76,77 but other studies have
been unable to link AD with this mood disorder.78,79 The
Metabolic syndrome potential mechanisms underlying the possible associa-
Instead of exploring the effect of its subcomponents, tion between these conditions might involve vascular
several studies have assessed the relationship between pathways and effects of depression on the hippocampal
metabolic syndrome as a whole and the risk of AD or formation or the hypothalamic–pituitary–adrenal axis.
cognitive decline. Most of these investigations demon-
strated a positive association between the presence of this Psychological stress
syndrome and cognitive dysfunction.62–64 Evidence from rodent studies suggests that chronic
psycho­logical stress can alter brain morphology (such
Smoking as hippocampal structure) and, as a result, exert a detri­
The relationship between smoking and cognitive decline mental effect on brain function, including memory.80
remains uncertain. Case–control studies have largely Thus, chronic psychological stress might increase the
suggested that smoking lowers the risk of AD, 65–67 risk of AD.
whereas prospective studies have shown that smoking
increases this risk68–70 or has no effect on the probability Traumatic brain injury
of deve­loping AD.71,72 A meta-analysis that examined the Retrospective studies 81–83 suggested that individuals
relationship between smoking and AD while accounting with a history of traumatic brain injury (TBI) had a
for tobacco-industry affiliation found that the combined higher risk of dementia than individuals with no history
results of 18 cross-sectional studies without industry of such injury. Two meta-analyses84,85 demonstrated
affiliations yielded no association.73 By contrast, data that among patients with TBI, the risk of dementia was
from eight cross-sectional studies with tobacco-industry higher in men than in women. Prospective studies of the
affiliations suggested that smoking protected against relationship between TBI and AD have proved incon-
AD. Analysis of 14 cohort studies without tobacco- sistent,86–88 but post­mortem and experimental studies
industry affiliations yielded a significant increase in the support a link between these conditions. Evidence also
risk of AD in smokers. exists that after human brain injury, the extent of Aβ
Smoking could affect the risk of AD via several pathology 89 and tau patho­logy increases in brain tissue,
mecha­nisms. Smoking may increase the generation of cerebrospinal fluid (CSF) Aβ levels are elevated, and
free radicals, leading to high oxidative stress, or affect APP is overproduced.90

142  |  MARCH 2011  |  VOLUME 7  www.nature.com/nrneurol


© 2011 Macmillan Publishers Limited. All rights reserved
REVIEWS

Protective factors Physical activity


Diet Epidemiological and experimental data suggest
Diets high in fish, fruit and vegetables are high in anti- that physical exercise may promote brain health.
oxidants and polyunsaturated fatty acids (PUFAs). In Conflicting results have, however, emerged from
some observational population-based studies, people cross-sectional and longitudinal observational studies
who had a high intake of vitamins E and C (both anti- that examined the relationship between exercise levels
oxidants) were less likely to show cognitive decline and and cognitive decline or dementia: while some studies
had a lower AD risk than individuals with a low intake indicated that physical activity has a beneficial effect
of these vitamins.91–93 By contrast, other large prospective on brain health, others showed no association between
studies of the effects of vitamins on AD risk found no such these variables.122–126
associations,94,95 and investigations examining the effect of Physical activity could affect cognition via multi­
dietary PUFAs on the risk of cognitive dysfunction proved ple mechanisms. An improvement in aerobic fitness
inconclusive. Indeed, while several studies showed that increases cerebral blood flow, oxygen extraction and
the consumption of PUFAs led to reductions in the risks glucose utilization,122 and activates growth factors that
of dementia and AD,96–98 MCI99 and age-related cognitive promote structural brain changes, such as an increase in
decline,100 other studies found no association between capillary density.128 In addition, rodent studies suggest
dietary PUFAs and cognitive impairment.101 Scarmeas that physical activity decreases the rate of amyloid
et al. reported that consumption of a Mediterranean-type plaque formation.129
diet (MeDi)—a diet characterized by a high intake of RCTs exploring the effects of exercise on cognitive func-
plant foods and fish (with olive oil as the primary source tion in healthy elderly individuals have yielded conflict-
of monounsaturated fat), a moderate intake of wine and ing results.130–133 A recent meta-analysis that included 11
a low intake of red meat and poultry—reduced the inci- RCTs involving cognitively healthy people aged >55 years
dence of AD102 and showed a trend towards reducing the suggested that undertaking of aerobic physical activities
risk of MCI.103 These effects were independent of levels of improves selective cognitive functions, including cognitive
physical activity 104 and vascular comorbidity.105 In a sub- speed, as well as auditory and visual attention.134
sequent cohort study in France, MeDi was found not to
alter performance on the Isaacs Set Test, the Benton Visual Intellectual activity
Retention Test or the Free and Cued Selective Reminding Following initial reports that elderly people with higher
Test, but was associated with high MMSE scores,106 pro- levels of education had a lower incidence of dementia
viding some support for the findings from the initial than individuals with no education, cognitive activity
studies of this diet. In a meta-analysis of 15 prospective was suggested to decrease the risk of cognitive decline by
studies exploring the effect of alcohol on dementia risk,107 increasing cognitive reserve. Several prospective studies
light to moderate alcohol consumption was associated subsequently found that both young 135 and old136 people
with a reduction in the risk of AD and dementia. who engage in cognitively stimulating activities, such as
Most RCTs examining the effects of antioxidant sup- learning, reading or playing games, were less likely to
plementation have found no association with cognitive develop dementia than individuals who did not engage
performance.108–111 To date, prospective clinical trial data in these activities.
for dietary supplementation with omega‑3 PUFAs have RCTs have shown a beneficial effect of intellectual
shown no overall effect on cognition in patients with interventions on cognitive function in elderly, dementia-
MCI or AD, but have suggested that docosahexaenoic ­free individuals.137 The benefits of cognitive training seem
acid supplementation has a beneficial effect on cognitive to be domain specific, however. Several trials found that
function in people harboring the APOE ε4 allele and in while cognitive training can improve memory, reasoning
the earliest stages of AD.112,113 and mental processing speed in older adults,137 cognitive
Reactive oxygen species are clearly associated with training did not have an effect on all cognitive domains,
neuronal damage in AD; however, whether the presence and did not affect day-to-day functioning.138 In addition,
of these molecules reflects a primary or secondary event one study found that among elderly individuals, those
in the neurotoxic process remains unclear. Deposition with memory impairment showed less improvement in
of Aβ, which is an early event in AD, leads to a decrease cognition through memory training than those without
in cerebral iron and copper concentrations, resulting in such impairment.139 Consequently, in elderly people, the
oxidative stress and neuronal damage.114 Evidence from effect of cognitive training on the risk of dementia is
in vitro studies indicates that vitamin E reduces the extent unclear, but several trials are underway.
of Aβ-induced lipid peroxidation and cell death.115 In
addition, carotenes and vitamin C protect against lipid Genetic epidemiology
peroxidation.116 Furthermore, vitamin C reduces the for- AD is usually classified according to its age of onset. The
mation of nitrosamines and may affect catecholamine majority (>95%) of patients who develop this disease are
synthesis. 117,118 Evidence also exists that antioxidant aged >65 years (so-called late-onset AD), with 1–5% of
intake reduces AD risk through a reduction in the risk AD cases exhibiting an earlier onset, typically in the late
of cerebro­vascular disease.119 Besides reducing oxidative 40s or early 50s (so-called early-onset AD). Late-onset
stress, PUFAs have favorable effects on neuronal and and early-onset AD are clinically indistinguishable;
vascular functions and inflammatory processes.120,121 however, the latter is generally more severe than the

NATURE REVIEWS | NEUROLOGY VOLUME 7  |  MARCH 2011  |  143


© 2011 Macmillan Publishers Limited. All rights reserved
REVIEWS

former and is associated with a more rapid rate of pro- of 58–79% for late-onset AD, depending on the model
gression. Moreover, the two forms of AD are associated that was used in the data analysis.
with different patterns of genetic epidemiology.
Apolipoprotein E
Early-onset Alzheimer disease APOE is the only established susceptibility gene for late-
Three genes have been firmly implicated in the onset AD and maps to chromosome 19 in a cluster with the
pathophysiology of early-onset AD, namely APP itself, genes encoding translocase of outer mitochondrial mem-
and the presenilin genes (PSEN1 and PSEN2), which brane 40 (TOMM40), apolipoprotein C1 and apolipo­
encode proteins involved in APP breakdown and Aβ gen- protein C2. APOE is a lipid-binding protein that is
eration. AD‑linked mutations in these three genes can be expressed in humans as one of three common isoforms,
considered to be ‘diagnostic biomarkers’ of this disease: which are encoded by three different alleles, namely APOE
these mutations exhibit high penetrance (>85%), mostly ε2, APOE ε3 and APOE ε4. The presence of a single
show autosomal dominant inheritance, and lead with APOE ε4 allele is associated with a 2–3-fold increase in the
certainty to Aβ aggregation and early-onset disease. risk of AD, while the presence of two copies of this allele is
APP mutations account for <0.1% of AD cases. Most associated with a fivefold increase in the risk of this disease.
dominantly inherited AD‑linked missense mutations in Each inherited APOE ε4 allele lowers the age of AD onset
APP affect the processing of the encoded protein, since by 6–7 years.145–147 Furthermore, the presence of this allele
the mutations are positioned in or near the Aβ-coding is associated with memory impairment, MCI, and progres-
exons (APP exons 16 and 17). 140 In addition to these sion from MCI to dementia.148 APOE ε4 has been suggested
dominant mutations, the APP mutation spectrum to account for as much as 20–30% of AD risk.
extends to two recessive mutations (which only cause Despite the studies linking APOE ε4 with AD, the
disease in the homozygous state), as well as APP locus presence of this allele is neither necessary nor sufficient
duplications, underscoring the importance of APP gene for disease: among participants in the Framingham
dosage in AD.140 study,149 55% of those who were homozygous for APOE
At present, 182 different AD‑related mutations from ε4, 27% of those with one copy of this allele and 9%
401 families have been identified in PSEN1, while of those without an APOE ε4 allele developed AD by
only 14 AD‑linked mutations from 23 families have 85 years of age. Segregation analyses conducted in fami-
been detected in PSEN2.140 The majority of AD‑linked lies of patients with AD support the presence of at least
PSEN1 and PSEN2 mutations are single-nucleotide four to six additional major AD risk genes.150
substitutions, but small deletions and insertions have
been described as well. The presenilins are functionally Additional genetic risk variants
involved in the γ‑secretase-mediated proteolytic cleav- After APOE, the best-validated gene modulating late-
age of APP.141 Mutations in PSEN1 and PSEN2 impair onset AD risk is the sortilin-related receptor 1 (SORL1)
this cleavage and cause an increase in the Aβ1–42:Aβ1–40 gene, which is located on chromosome 11q23. SorL1
ratio. This rise might occur through either an increase belongs to a group of five type I transmembrane receptors
in Aβ1–42 levels, as indicated in plasma and fibroblast (the others being sortilin, SorCS1, SorCS2 and SorCS3)
media from PSEN-mutation carriers,142 or a decrease in that are highly expressed in the CNS and are character-
Aβ1–40 levels, suggesting a loss-of-function rather than a ized by a luminal, extracellular vacuolar protein sorting 10
gain-of-function mechanism. domain. From family-based and population-based studies
To summarize, all three causal AD genes lend support that, together, included over 6,000 individuals from four
to a common pathogenic AD pathway, with a pivotal role ethnic groups, Rogaeva et al. identified two haplotypes
for Aβ. According to this amyloid hypothesis, neuro­ in the 3' and 5' regions of SORL1 that are associated
degenerative processes in AD are the consequence of with late-onset AD risk.151 In addition, these research-
an imbalance between Aβ production and Aβ clear- ers demonstrated that SorL1 promotes the translocation
ance, suggesting that other genes involved in these and retention of APP in subcellular compartments that
pathways might also be risk factors for this disease. exhibit low secretase activity, thereby reducing the extent
of proteolytic breakdown into both amyloidogenic and
Late-onset Alzheimer disease nonamyloidogenic products.151 As a consequence, under-
The genes involved in late-onset AD increase disease expression of SORL1 leads to overexpression of Aβ and an
risk and are not inherited in a Mendelian fashion. First- increased risk of AD. Several subsequent studies replicated
degree relatives of patients with late-onset AD have these initial genetic association findings, and the results
twice the expected lifetime risk of this disease of people were further validated by a collaborative, unbiased meta-
who do not have an AD‑affected first-degree relative.143 analysis of all published genetic data sets that included a
In addition, AD occurs more frequently in monozygotic total of 12,464 AD cases and 17,929 controls.152 In the past
than in dizygotic co-twins,144 suggesting a substantial year, two studies demonstrated that, in addition, genetic
genetic contribution to this disorder. In the largest variation in the SORL1 homolog SORCS1 influences
twin study of dementia, involving 11,884 participants AD risk, cognitive performance, APP processing and
in the Swedish registry who were aged >65 years, 395 Aβ1–40 and Aβ1–42 levels through an effect on γ‑secretase
twin pairs were identified in which either one or both processing of APP,153,154 further emphazising the role of
twins had AD.144 This study demonstrated a heritability SorL1-related proteins in late-onset AD etiology.

144  |  MARCH 2011  |  VOLUME 7  www.nature.com/nrneurol


© 2011 Macmillan Publishers Limited. All rights reserved
REVIEWS

Genome-wide association studies for AD using Box 4 | Endophenotypes and genetic epidemiology
large numbers of cases and controls155–158 have revealed
modest effect sizes for several genes on AD risk, with Endophenotypes (measurable intermediate phenotypes that are closer to the action
odds ratios in the range of 1.1–1.5, although most of of the gene than affection status) provide clues to the genetic underpinnings of a
disease, and diagnoses can be deconstructed to increase the success of genetic
these studies have only confirmed the association
analyses. Age at disease onset is a well-established endophenotype for Alzheimer
of APOE with this disease. One such study showed that disease (AD), and is a key indicator of genetic heterogeneity. Recognition of
variants of TOMM40—which is proximally located to various age-at-onset profiles among families was critical in the initial identification
and in linkage disequilibrium with APOE—were associ- of genetically distinct forms of AD. Age information, such as age-at-onset data
ated with AD risk, but whether these genetic associations for individuals affected by AD and censored age data for unaffected individuals,
are independent of the APOE locus remains unclear.159 is also an important covariate used to modify the penetrance function, allowing
Together, two genome-wide association studies determination of age-dependent penetrance of the disease as a function of disease
genotype, while also increasing the power of genetic linkage studies and the
identified variants in the clusterin gene (CLU), the
accuracy with which disease genes can be localized. In addition to the use of age at
phosphatidylinositol-binding clathrin assembly protein onset as a covariate and as a stratifier in genome scans, this endophenotype has
gene (PICALM) and complement receptor type 1 gene a genetic basis, with several contributing loci having been identified. Nevertheless,
(CR1) as being associated with AD, but functional data despite overwhelming evidence that age is an important variable to be considered
confirming the roles in AD of the proteins encoded by in genetic risk factor research for AD, most genome scans for this disease have not
these genes are still lacking.160,161 Clusterin is a lipoprotein incorporated age information. Other quantitative phenotypes that have been proven
that is expressed in mammalian tissues and is incorporated useful are pathological phenotypes, such as neuritic plaque and neurofibrillary
into amyloid plaques. This protein binds to soluble Aβ in tangle densities, structural brain changes on brain imaging such as white matter
lesions, and cognitive function.
CSF, forming complexes that can penetrate the BBB.
Clusterin levels are positively correlated with the number
of APOE ε4 alleles, suggesting a compensatory induction Box 5 | Epigenetic mechanisms
of CLU in the brains of AD patients with the APOE ε4
allele, who show low brain levels of APOE.162 Several epigenetic studies are underway using normal
brain tissue to identify functional cis-acting regulatory
CR1 encodes a protein that is likely to contribute to
polymorphisms that may be associated with brain
Aβ clearance from the brain,163 while PICALM protein is disorders such as Alzheimer disease. Allele-specific DNA
involved in clathrin-mediated endocytosis, allowing intra- methylation—an epigenetic marker—is often strongly
cellular trafficking of proteins and lipids such as nutrients, dependent on nearby single-nucleotide polymorphisms
growth factors and neurotransmitters.164 PICALM protein (SNPs) and haplotypes. This observation suggests that
also has a role in the trafficking of vesicle-associated mem- mapping allele-specific DNA methylation across the whole
brane protein 2, a soluble N‑ethylmaleimide-sensitive genome in human brain DNA could be useful for finding
factor attachment protein receptor that is involved in the regulatory SNPs that act in this target tissue and are
involved in specific brain disorders.
fusion of synaptic vesi­cles to the presynaptic membrane
in neurotransmitter release.
A third large genome-wide association study con- an observed genetic association. Furthermore, while
firmed the associations of PICALM and CLU with AD165 genome-wide association studies represent a method of
and reported two additional loci as being associated with screening the genome, limitations exist in their ability
AD: rs744373, which is near the bridging inte­grator 1 to detect true associations. Also, the results of such
gene (BIN1) on chromosome 2q14.3, and rs597668, studies can be difficult to replicate if the actual effect is
which is located on chromosome 19q13.3. smaller than that observed in the initial study. Finally,
BIN1 is a member of the BAR (BIN–amphiphysin–Rvs) the detection of associations with multiple rare variants
adaptor family, which has been implicated in caspase- at a single site (which are better detected by linkage
independent apoptosis and membrane dynamics, includ- studies) or with single rare variants (minor allele fre-
ing vesicle fusion and trafficking, neuronal membrane quency <5%) may not be possible. These limitations
organization, and clathrin-mediated sy­naptic vesicle for- have led researchers to focus on additional biological
mation.166 Of note, the latter process is disrupted by Aβ.167 characteristics—including endo­phenotypes (Box 4)
Changes in BIN1 expression have also been shown in aging and epigenetic characteristics (Box 5)—to facilitate the
mice and in transgenic mouse models of AD.168 identification of disease-causing mutations.
The locus rs597668 is not in linkage disequilibrium
with APOE, suggesting that the effect of this locus on AD Biomarkers
risk is independent. Six genes are found in this region, The epidemiological research exploring environmental
of which at least two (genes encoding biogenesis of lyso- and genetic risk factors for late-onset AD has led to the
somal organelles complex 1 subunit 3 and microtubule- identification of various biomarkers for this disease. In
associated protein–microtubule affinity-regulating addition to being helpful tools for the determination of
kinase 4) are implicated in molecular pathways linked to disease risk, biomarkers are invaluable in establishing a
AD or other brain disorders.169–171 diagnosis. In contrast to the autosomal dominantly inher-
The results of the published genome-wide associ­ation ited mutations reviewed above, the additional biomarkers
studies are informative, but the genetic associations need that have been identified for AD are not definite markers of
functional validation. Indeed, such studies alone cannot disease but, rather, contribute to increasing the specificity
prove causality or assess the biological significance of of diagnosis. These additional biomarkers include various

NATURE REVIEWS | NEUROLOGY VOLUME 7  |  MARCH 2011  |  145


© 2011 Macmillan Publishers Limited. All rights reserved
REVIEWS

CSF APOE levels correlate positively with CSF levels of


t‑tau and 24S-hydroxycholesterol in patients with cogni-
tive disorders.181 Supplementary Table 1 shows additional
CSF biomarkers that have been reported for predicting
the risk or aiding the diagnosis of AD. In a large-scale
meta-analysis of 14 studies exploring biomarkers for
pre­clinical AD, the effect sizes of Aβ1–42, t‑tau and p‑tau
ranged from 0.91–1.11, suggesting that the sensitivity of
these markers for this condition is modest.182

Plasma biomarkers
Owing to the properties of the BBB, plasma bio­markers
comprise only small or lipophilic proteins, and proteins
carried by specific transporters that are able to cross this
barrier. The exact mechanisms underlying the inter­
action between brain and plasma Aβ levels remain
unclear; however, under physiological conditions, a
steady-state level of brain Aβ exists that is maintained
Figure 2 | T1-weighted MRI scan of a patient with a clinical
diagnosis of late-onset AD. As is typical for late-onset AD,
by a balance between the production and deposition of
the MRI scan shows generalized brain atrophy and loss of Aβ in the brain and the production of this peptide in
gray matter affecting the hippocampus (red arrow), the periphery by platelets. RAGEs transport plasma Aβ
entorhinal cortex (green arrow) and perirhinal cortex through the BBB into the brain, whereas the cell-surface
(blue arrow). Abbreviation: AD, Alzheimer disease. low-density lipoprotein receptor related protein‑1 trans-
ports this peptide from the brain into plasma. In healthy
individuals, brain Aβ levels are, therefore, reflected by
measurements from CSF, blood and neuro­imaging. The plasma Aβ concentrations. In patients with dementia, the
reliability, specificity and sensitivity of these biomarkers relationship between brain Aβ levels and the concentra-
are determined by epidemiological studies. tion of Aβ in plasma is unclear, owing to deposition of
Aβ in amyloid plaques.
Cerebrospinal fluid biomarkers In familial AD 142,183 and Down syndrome with
The free transport of proteins between the brain and CSF APP triplication,184 total Aβ levels and Aβ1–42 levels in
provides a reflection of the cerebral metabolic processes plasma are elevated. In sporadic AD, the usefulness of
occurring in the latter. CSF levels of Aβ1–42, total tau (t-tau) plasma Aβ as a risk biomarker remains controversial.185,186
and p‑tau are validated markers that may aid the accurate Studies found associations between plasma Aβ1–40 levels
diagnosis of AD at an early stage of disease.172 CSF samples and AD,186 reported no association between AD risk
from patients diagnosed as having AD or MCI contain and plasma Aβ levels,187 showed that plasma Aβ1–42 levels
lower levels of Aβ1–42 and higher levels of t‑tau or p‑tau are reduced in patients with AD compared with healthy
than do those from cognitively normal individuals.173 controls,188 or demonstrated that levels of plasma Aβ1–42
Furthermore, t‑tau CSF levels increase with AD progres- decrease before the onset of AD.189 These contrasting
sion.174 A longitudinal study that aimed to clarify whether findings result from variability in the timing of sample
variation in the concentration pattern of different-length collection across studies, the use of different antibodies
Aβ peptides increases diagnostic power suggested that to detect Aβ, and a lack of validation of plasma Aβ as a
the combined levels of Aβ 1–37, Aβ 1–38, Aβ 1–39, Aβ 1–40 risk biomarker for AD.
and Aβ1–42 have a 91% sensitivity and a 64% specificity One study showed that in comparison with individu-
in predicting progression from MCI to AD.175 als who remained healthy during the follow-up period,
The relationship between concentrations of amyloid participants who later went on to develop AD showed a
plaques and NFTs in the brain and levels of CSF bio­ significant increase in plasma Aβ1–42 levels but no rise in
markers (that is, Aβ1–42, t‑tau and p‑tau) remains to be plasma Aβ1–40 levels after the initial examination.189 At
clarified. According to one hypothesis,176 accumula- the time of conversion to AD, however, levels of plasma
tion of protofibrillar Aβ species in the brain may result Aβ1–42 and the Aβ1–42:Aβ1–40 ratio decreased significantly.
in a reduction of soluble Aβ in the CSF and brain, and These findings indicate that elevated plasma Aβ1–42 is an
increased Aβ deposition in amyloid plaques. Some studies antecedent risk factor for AD, while decreasing levels of
suggested a positive correlation between p‑tau CSF levels plasma Aβ1–42 or a decline in the Aβ1–42:Aβ1–40 ratio mark
and NFT concentration in people with AD;177 however, disease onset.
other studies were unable to validate these findings.178,179 Several other molecules have been investigated as
In the interpretation of the data generated in these studies, plasma biomarkers for AD risk, including homocysteine,
the importance of various influencing factors on Aβ and various proteins linked to inflammation (for example,
tau levels in CSF needs to be recognized. Increasing age C‑reactive protein, IL‑1β, TNF, IL‑6 and transforming
or carrying the APOE ε4 allele accelerates the deposition growth factor β), and cholesterol. These molecules have,
of Aβ1–42 in the brain and lowers CSF Aβ1–42 levels.180Also, however, all yielded inconsistent data across studies.

146  |  MARCH 2011  |  VOLUME 7  www.nature.com/nrneurol


© 2011 Macmillan Publishers Limited. All rights reserved
REVIEWS

a 18 b 11
F-FDG C-PIB
Low uptake High uptake Low retention High retention

Normal aging Late-onset AD Normal aging Late-onset AD

Figure 3 | Changes revealed by PET in the AD brain. a | 18F-FDG-PET patterns characteristic of metabolic activity in
cognitively normal individuals and patients with late-onset AD. In comparison with people aging normally, individuals with
late-onset AD show decreased bilateral glucose metabolism, particularly in the temporal and parietal regions. b | 11C-PIB
PET images characteristic of elderly individuals without cognitive impairment and patients with late-onset AD. The high
concentrations of 11C-PIB in the AD brain are suggestive of high amounts of amyloid deposits. Abbreviations: AD,
Alzheimer disease; FDG, 2‑fluoro‑2-deoxy‑d-glucose; PIB, Pittsburgh compound B.

Structural MRI fMRI, which measures alterations in blood flow on


Late-onset AD is characterized by medial temporal the basis of changes in deoxyhemoglobin. As the
lobe atrophy, particularly in the hippocampal forma- deoxyhemo­globin concentration depends on neuronal
tion and the amygdala (Figure 2). In early-onset AD, activity, BOLD reflects brain activity. This technique is
brain atrophy may be located more posteriorly than in widely used in research and in the diagnosis of various
late-onset AD, and may involve the posterior cortex,190 brain disorders because of its high sensitivity and
occipital lobes, precuneus, posterior cingulate191 and easy implementation.
amygdala. 192Atrophy in the hippocampus and ento- BOLD signals depend on several anatomical, physio­
rhinal cortex is associated with a decline in memory logical and imaging parameters, and can be interpreted
function, progression of memory impairment 193 and an qualitatively or semiquantitatively. As a result, inter­
increased risk of AD.194 These changes are not specific individual and intraindividual variability limits the
to AD, however, and substantial anatomical overlap use of such signals in the differential diagnosis of
prevails between the types of atrophy observed in AD, dementia-­causing disorders. Nevertheless, fMRI can
various other neurodegenerative disorders and normal facilitate the characterization of functional abnormali-
aging. Thus, such changes are not sufficient to estab- ties in specific diseases. People with AD exhibit reduced
lish a definitive diagnosis of AD. In addition, while brain activity in parietal and hippocampal regions in
non­specific white matter changes appear frequently comparison with healthy controls.197 In addition, some
in healthy elderly individuals, such changes are also studies have found different neuronal activity patterns
common in elderly people with cognitive decline, stroke in healthy controls and patients with MCI. 190 Recent
or MCI. Nevertheless, several studies have suggested advances in fMRI have allowed intrinsic functional
that certain structural MRI bio­markers possess some networks in the human brain to be defined. The study
degree of discriminative diagnostic power. Evidence of cognitive–behavioral function in the early stages of
exists that in AD, the corpus callosum (particularly neurodegenerative disorders may allow the identifica-
the anterior area) exhibits atrophy. This change helps tion of the neuroanatomical networks affected by these
to distinguish AD from fronto­temporal dementia, in diseases, and may assist in the differential diagnosis of
which the posterior area of the corpus callosum shows the various disorders that underlie dementia.
greater atrophy than the anterior area of this brain
structure. 195 Evidence is also available that among PET and single-photon emission CT
patients with amnestic MCI, those who convert to PET and single-photon emission CT (SPECT) have
AD show greater atrophy in the hippocampus and the been widely explored as diagnostic tools for dementia,
inferi­or and middle temporal gyri than those who do and both techniques have shown good diagnostic and
not convert to AD.196 prognostic capabilities. PET studies have mostly used
the tracer 2‑[18F]-fluoro‑2-deoxy‑d-glucose (18F-FDG),
Functional MRI which provides a measure of cerebral glucose metabo-
Functional MRI (fMRI) can visualize neuronal activ- lism and, hence, indirectly demonstrates synaptic activ-
ity either during rest or in association with a task that ity. In the early stages of AD, 18F-FDG-PET reveals a
activates specific brain regions. The most common characteristic pattern of symmetric hypo­metabolism
method is blood oxygen level-dependent (BOLD) in the pos­terior cingulate and parieto­temporal regions

NATURE REVIEWS | NEUROLOGY VOLUME 7  |  MARCH 2011  |  147


© 2011 Macmillan Publishers Limited. All rights reserved
REVIEWS

that spreads to the prefrontal cortices (Figure 3). These Conclusions


changes are distinct from the changes in cerebral Substantial progress has been made over the past few
glucose metabolism that are seen in healthy controls decades in understanding AD. Neverthless, many research-
and cases of other forms of dementia, and the extent of ers would argue that our knowledge of this disease is still
hypometabolism inversely correlates with the degree profoundly imperfect, as demonstrated by the failure of
of cognitive impairment. 198 18 F-FDG-PET has a all but symptomatic treatments for clinically diagnosed
high sensitivity (94%) but a low specificity (73– AD. We know that in people aged >85 years, dementia and
78%) for the diagnosis of dementia. 199 SPECT, cognitive impairment are common, reaching a combined
which involves studying regional blood flow with prevalence >50% in the oldest old, and that the incidence
Tc-hexamethylpropyleneamine oxime, has a similar of dementia continues to rise in the oldest age groups. Thus,
specificity to 18F-FDG-PET for this condition.200 screening is essential to identify cognitively normal individ-
A number of low-molecular-weight tracers have been uals in midlife or old age who have a high risk of developing
developed for PET to assess Aβ deposits in vivo. The MCI and AD, so that interventions, when available, can be
most frequently used tracer is Pittsburgh compound B administered to stop the development of specific disease-
(PIB). Compared with healthy controls, patients with related pathologies. In the oldest patients who develop AD,
AD show increased 11C-PIB retention in cortical regions for whom this disease is closely associated with end-of-life
targeted by Aβ deposits (Figure 3).201 Deposition of this issues, palli­ative therapies, rather than interventional treat-
peptide seems to reach a plateau by the early stages of ments, are required to ensure the best possible quality of life
AD.202 In MCI, PIB binding is bimodal, with ≈ 50% of and not necessarily an extension of lifespan.
patients showing an increase in 11C-PIB binding, resem-
bling the 11C-PIB retention that is seen in AD, while the
Review criteria
other ≈ 50% of patients exhibit low levels of 11C-PIB
binding that are similar to the levels seen in controls.203 Alzheimer disease genetic studies were identified
In MRI studies, 11C-PIB binding correlated positively from searches of the AlzGene database, which was
with atrophy in the amygdala and hippocampus but not updated in May 2010. PubMed was searched for articles
other cortical areas, suggesting that various brain areas published in English between January 1990 and October
2010. Combinations of the following key words were
have different susceptibilities to Aβ deposit-mediated
used in the aforementioned searches: “dementia”,
toxicity, or that amyloid deposition is nonessential for “Alzheimer’s disease”, “mild cognitive impairment”,
neurodegeneration.203 New PET tracers for amyloid “MCI”, “biomarker”, “risk”, “antecedent”, “gene”,
deposits, such as 18F-FDDNP, are being developed. In “genetics”, “epigenetics”, “endophenotype”, “incidence”,
studies comparing 11C-PIB and 18F-FDDNP, these tracers “prevalence”, “criteria”, “diagnosis, “definition”, “history”,
showed differences in regional binding and in the cogni- “pathology” and “autopsy”. The retrieved abstracts were
tive domains with which they seem to be associated, sug- read to identify studies addressing the topics included
in this Review. We also performed a manual search of
gesting that these tracers measure related but different
references cited in the published articles.
characteristics of AD.204

1. Alzheimer’s Association. 2010 Alzheimer’s 8. Kivipelto, M. et al. Midlife vascular risk factors 16. Kalaria, R. N. Vascular basis for brain
disease facts and figures. Alzheimers Dement. 6, and late-life mild cognitive impairment: degeneration: faltering controls and risk factors
158–194 (2010). a population-based study. Neurology 56, for dementia. Nutr. Rev. 68 (Suppl. 2), S74–S87
2. McKhann, G. et al. Clinical diagnosis of 1683–1689 (2001). (2010).
Alzheimer’s disease: report of the NINCDS– 9. Launer, L. J., Masaki, K., Petrovitch, H., Foley, D. 17. Deane, R., Wu, Z. & Zlokovic, B. V. RAGE (yin)
ADRDA Work Group under the auspices of & Havlik, R. J. The association between midlife versus LRP (yang) balance regulates Alzheimer
Department of Health and Human Services Task blood pressure levels and late-life cognitive amyloid β‑peptide clearance through transport
Force on Alzheimer’s Disease. Neurology 34, function. The Honolulu–Asia Aging Study. JAMA across the blood–brain barrier. Stroke 35,
939–944 (1984). 274, 1846–1851 (1995). 2628–2631 (2004).
3. Ferri, C. P. et al. Global prevalence of dementia: a 10. Swan, G. E., Carmelli, D. & Larue, A. Systolic 18. Forette, F. et al. The prevention of dementia with
Delphi consensus study. Lancet 366, blood pressure tracking over 25 to 30 years and antihypertensive treatment: new evidence from
2112–2117 (2005). cognitive performance in older adults. Stroke 29, the Systolic Hypertension in Europe (Syst-Eur.)
4. Pendlebury, S. T. & Rothwell, P. M. Prevalence, 2334–2340 (1998). study. Arch. Intern. Med. 162, 2046–2052
incidence, and factors associated with pre-stroke 11. Whitmer, R. A., Sidney, S., Selby, J., (2002).
and post-stroke dementia: a systematic review Johnston, S. C. & Yaffe, K. Midlife cardiovascular 19. Tzourio, C. et al. Effects of blood pressure
and meta-analysis. Lancet Neurol. 8, 1006–1018 risk factors and risk of dementia in late life. lowering with perindopril and indapamide
(2009). Neurology 64, 277–281 (2005). therapy on dementia and cognitive decline in
5. Cheung, Z. H., Gong, K. & Ip, N. Y. Cyclin- 12. Glynn, R. J. et al. Current and remote blood patients with cerebrovascular disease. Arch.
dependent kinase 5 supports neuronal survival pressure and cognitive decline. JAMA 281, Intern. Med. 163, 1069–1075 (2003).
through phosphorylation of Bcl‑2. J. Neurosci. 28, 438–445 (1999). 20. Starr, J. M., Whalley, L. J. & Deary, I. J. The
4872–4877 (2008). 13. Knopman, D. et al. Cardiovascular risk factors effects of antihypertensive treatment on
6. Weishaupt, J. H. et al. Inhibition of CDK5 is and cognitive decline in middle-aged adults. cognitive function: results from the HOPE study.
protective in necrotic and apoptotic paradigms of Neurology 56, 42–48 (2001). J. Am. Geriatr. Soc. 44, 411–415 (1996).
neuronal cell death and prevents mitochondrial 14. Posner, H. B. et al. The relationship of 21. Prince, M. J., Bird, A. S., Blizard, R. A.
dysfunction. Mol. Cell. Neurosci. 24, 489–502 hypertension in the elderly to AD, vascular & Mann, A. H. Is the cognitive function of older
(2003). dementia, and cognitive function. Neurology 58, patients affected by antihypertensive
7. Wen, Y. et al. Cdk5 is involved in NFT-like 1175–1181 (2002). treatment? Results from 54 months of the
tauopathy induced by transient cerebral ischemia 15. Skoog, I. et al. 15-year longitudinal study of Medical Research Council’s trial of
in female rats. Biochim. Biophys. Acta 1772, blood pressure and dementia. Lancet 347, hypertension in older adults. BMJ 312,
473–483 (2007). 1141–1145 (1996). 801–805 (1996).

148  |  MARCH 2011  |  VOLUME 7  www.nature.com/nrneurol


© 2011 Macmillan Publishers Limited. All rights reserved
REVIEWS

22. [No authors listed] Prevention of stroke by 40. Watson, G. S. et al. Preserved cognition in 59. Simons, M. et al. Treatment with simvastatin in
antihypertensive drug treatment in older patients with early Alzheimer disease and normocholesterolemic patients with Alzheimer’s
persons with isolated systolic hypertension. amnestic mild cognitive impairment during disease: a 26-week randomized, placebo-
Final results of the Systolic Hypertension in the treatment with rosiglitazone: a preliminary study. controlled, double-blind trial. Ann. Neurol. 52,
Elderly Program (SHEP). SHEP Cooperative Am. J. Geriatr. Psychiatry 13, 950–958 (2005). 346–350 (2002).
Research Group. JAMA 265, 3255–3264 41. Risner, M. E. et al. Efficacy of rosiglitazone in a 60. Sparks, D. L. et al. Circulating cholesterol levels,
(1991). genetically defined population with apolipoprotein E genotype and dementia
23. Lithell, H. et al. The Study on Cognition and mild‑to‑moderate Alzheimer’s disease. severity influence the benefit of atorvastatin
Prognosis in the Elderly (SCOPE): principal Pharmacogenomics J. 6, 246–254 (2006). treatment in Alzheimer’s disease: results of the
results of a randomized double-blind 42. Sato, T. et al. Efficacy of PPAR‑γ agonist Alzheimer’s Disease Cholesterol-Lowering
intervention trial. J. Hypertens. 21, 875–886 pioglitazone in mild Alzheimer disease. Treatment (ADCLT) trial. Acta Neurol. Scand.
(2003). Neurobiol. Aging doi:10.1016/ Suppl. 185, 3–7 (2006).
24. Peters, R. et al. Association of depression with j.neurobiolaging.2009.10.009. 61. Sano, M. Multi-center, randomized, double-blind,
subsequent mortality, cardiovascular morbidity 43. Jiang, Q., Heneka, M. & Landreth, G. E. The role placebo-controlled trial of simvatatin to slow the
and incident dementia in people aged 80 and of peroxisome proliferator-activated receptor- progression of Alzheimer’s disease. Alzheimers
over and suffering from hypertension. Data from gamma (PPARγ) in Alzheimer’s disease: Dement. 4 (Suppl. 2), T200 (2008).
the Hypertension in the Very Elderly Trial therapeutic implications. CNS Drugs 22, 1–14 62. Raffaitin, C. et al. Metabolic syndrome and risk
(HYVET). Age Ageing 39, 439–445 (2010). (2008). for incident Alzheimer’s disease or vascular
25. Leibson, C. L. et al. The risk of dementia among 44. Grundman, M., Corey-Bloom, J., Jernigan, T., dementia: the Three-City Study. Diabetes Care
persons with diabetes mellitus: a population- Archibald, S. & Thal, L. J. Low body weight in 32, 169–174 (2009).
based cohort study. Ann. N. Y. Acad. Sci. 826, Alzheimer’s disease is associated with mesial 63. Solfrizzi, V. et al. Metabolic syndrome and the
422–427 (1997). temporal cortex atrophy. Neurology 46, risk of vascular dementia: the Italian
26. Luchsinger, J. A., Tang, M. X., Stern, Y., Shea, S. 1585–1591 (1996). Longitudinal Study on Ageing. J. Neurol.
& Mayeux, R. Diabetes mellitus and risk of 45. White, H., Pieper, C. & Schmader, K. The Neurosurg. Psychiatry 81, 433–440 (2010).
Alzheimer’s disease and dementia with stroke in association of weight change in Alzheimer’s 64. Yaffe, K., Weston, A. L., Blackwell, T.
a multiethnic cohort. Am. J. Epidemiol. 154, disease with severity of disease and mortality: & Krueger, K. A. The metabolic syndrome and
635–641 (2001). a longitudinal analysis. J. Am. Geriatr. Soc. 46, development of cognitive impairment among
27. Ott, A. et al. Diabetes mellitus and the risk of 1223–1227 (1998). older women. Arch. Neurol. 66, 324–328
dementia: The Rotterdam Study. Neurology 53, 46. Gustafson, D., Rothenberg, E., Blennow, K., (2009).
1937–1942 (1999). Steen, B. & Skoog, I. An 18-year follow-up of 65. Tyas, S. L. Are tobacco and alcohol use related
28. Craft, S. Insulin resistance and Alzheimer’s overweight and risk of Alzheimer disease. Arch. to Alzheimer’s disease? A critical assessment
disease pathogenesis: potential mechanisms Intern. Med. 163, 1524–1528 (2003). of the evidence and its implications. Addict. Biol.
and implications for treatment. Curr. Alzheimer 47. Razay, G. & Vreugdenhil, A. Obesity in middle 1, 237–254 (1996).
Res. 4, 147–152 (2007). age and future risk of dementia: midlife obesity 66. Brenner, D. E. et al. Relationship between
29. Cook, D. G. et al. Reduced hippocampal insulin- increases risk of future dementia. BMJ 331, cigarette smoking and Alzheimer’s disease in a
degrading enzyme in late-onset Alzheimer’s 455 (2005). population-based case–control study. Neurology
disease is associated with the apolipoprotein 48. Stewart, R. et al. A 32-year prospective study of 43, 293–300 (1993).
E‑ε4 allele. Am. J. Pathol. 162, 313–319 (2003). change in body weight and incident dementia: 67. Ferini-Strambi, L., Smirne, S., Garancini, P.,
30. Yamagishi, S., Nakamura, K., Inoue, H., the Honolulu–Asia Aging Study. Arch. Neurol. Pinto, P. & Franceschi, M. Clinical and
Kikuchi, S. & Takeuchi, M. Serum or 62, 55–60 (2005). epidemiological aspects of Alzheimer’s disease
cerebrospinal fluid levels of glyceraldehyde- 49. Gustafson, D. R. et al. Adiposity indicators and with presenile onset: a case control study.
derived advanced glycation end products (AGEs) dementia over 32 years in Sweden. Neurology Neuroepidemiology 9, 39–49 (1990).
may be a promising biomarker for early 73, 1559–1566 (2009). 68. Merchant, C. et al. The influence of smoking on
detection of Alzheimer’s disease. Med. 50. Whitmer, R. A. et al. Central obesity and the risk of Alzheimer’s disease. Neurology 52,
Hypotheses 64, 1205–1207 (2005). increased risk of dementia more than three 1408–1412 (1999).
31. Yan, S. D. et al. RAGE and amyloid‑β peptide decades later. Neurology 71, 1057–1064 69. Launer, L. J. et al. Rates and risk factors for
neurotoxicity in Alzheimer’s disease. Nature (2008). dementia and Alzheimer’s disease: results from
382, 685–691 (1996). 51. Muckle, T. J. & Roy., J. R. High-density EURODEM pooled analyses. EURODEM
32. Harvey, J., Solovyova, N. & Irving, A. Leptin and lipoprotein cholesterol in differential diagnosis Incidence Research Group and Work Groups.
its role in hippocampal synaptic plasticity. Prog. of senile dementia. Lancet 1, 1191–1193 European Studies of Dementia. Neurology 52,
Lipid Res. 45, 369–378 (2006). (1985). 78–84 (1999).
33. Li, X. L. et al. Impairment of long-term 52. Kuo, Y. M. et al. Elevated low-density lipoprotein 70. Ott, A. et al. Smoking and risk of dementia and
potentiation and spatial memory in leptin in Alzheimer’s disease correlates with brain Alzheimer’s disease in a population-based
receptor-deficient rodents. Neuroscience 113, Aβ 1–42 levels. Biochem. Biophys. Res. cohort study: the Rotterdam Study. Lancet 351,
607–615 (2002). Commun. 252, 711–715 (1998). 1840–1843 (1998).
34. Lieb, W. et al. Association of plasma leptin levels 53. Michikawa, M. Cholesterol paradox: is high 71. Doll, R., Peto, R., Boreham, J. & Sutherland, I.
with incident Alzheimer disease and MRI total or low HDL cholesterol level a risk for Smoking and dementia in male British doctors:
measures of brain aging. JAMA 302, Alzheimer’s disease? J. Neurosci. Res. 72, prospective study. BMJ 320, 1097–1102
2565–2572 (2009). 141–146 (2003). (2000).
35. Profenno, L. A, Porsteinsson, A. P. & 54. Wieringa, G. E. et al. Apolipoprotein E genotypes 72. Hebert, L. E. et al. Relation of smoking and
Faraone, S. V. Meta-analysis of Alzheimer’s and serum lipid levels in Alzheimer’s disease alcohol consumption to incident Alzheimer’s
disease risk with obesity, diabetes, and related and multi-infarct dementia. Int. J. Geriatr. disease. Am. J. Epidemiol. 135, 347–355
disorders. Biol. Psychiatry 67, 505–512 (2010). Psychiatry 12, 359–362 (1997). (1992).
36. Janson, J. et al. Increased risk of type 2 55. van Exel, E. et al. Association between high- 73. Cataldo, J. K., Prochaska, J. J. & Glantz, S. A.
diabetes in Alzheimer disease. Diabetes 53, density lipoprotein and cognitive impairment in Cigarette smoking is a risk factor for Alzheimer’s
474–481 (2004). the oldest old. Ann. Neurol. 51, 716–721 disease: an analysis controlling for tobacco
37. Peila, R., Rodriguez, B. L. & Launer, L. J. Type 2 (2002). industry affiliation. J. Alzheimers Dis. 19,
diabetes, APOE gene, and the risk for dementia 56. Lesser, G. et al. Elevated serum total and LDL 465–480 (2010).
and related pathologies: the Honolulu–Asia cholesterol in very old patients with Alzheimer’s 74. Traber, M. G., van der Vliet, A., Reznick, A. Z.
Aging Study. Diabetes 51, 1256–1262 (2002). disease. Dement. Geriatr. Cogn. Disord. 12, & Cross, C. E. Tobacco-related diseases. Is
38. Arvanitakis, Z. et al. Diabetes is related to 138–145 (2001). there a role for antioxidant micronutrient
cerebral infarction but not to AD pathology in 57. Burns, M. & Duff, K. Cholesterol in Alzheimer’s supplementation? Clin. Chest Med. 21,
older persons. Neurology 67, 1960–1965 disease and tauopathy. Ann. N. Y. Acad. Sci. 173–187 (2000).
(2006). 977, 367–375 (2002). 75. Kellar, K. J. & Wonnacott, S. in Nicotine
39. Reger, M. A. et al. Effects of intranasal insulin 58. Jones, R. W. et al. The Atorvastatin/Donepezil Psychopharmacology: Molecular, Cellular,
on cognition in memory-impaired older adults: in Alzheimer’s Disease Study (LEADe): design and Behavioral Aspects (eds Wonnacott, S.,
modulation by APOE genotype. Neurobiol. Aging and baseline characteristics. Alzheimers Russell, M. A. & Stolerman, I. P) 341–373
27, 451–458 (2006). Dement. 4, 145–153 (2008). (Oxford University Press, Oxford, 1990).

NATURE REVIEWS | NEUROLOGY VOLUME 7  |  MARCH 2011  |  149


© 2011 Macmillan Publishers Limited. All rights reserved
REVIEWS

76. Jorm, A. F. History of depression as a risk intake of antioxidants and risk of late-life disease: the OmegAD study. Dement. Geriatr.
factor for dementia: an updated review. Aust. incident dementia: the Honolulu–Asia Aging Cogn. Disord. 27, 481–490 (2009).
NZ J. Psychiatry 35, 776–781 (2001). Study. Am. J. Epidemiol. 159, 959–967 (2004). 114. Nagano, S. et al. Peroxidase activity of
77. Barnes, D. E., Alexopoulos, G. S., Lopez, O. L., 95. Luchsinger, J. A., Tang, M. X., Shea, S. cyclooxygenase‑2 (COX‑2) cross-links β‑amyloid
Williamson, J. D. & Yaffe, K. Depressive & Mayeux, R. Antioxidant vitamin intake and risk (Aβ) and generates Aβ–COX‑2 hetero-oligomers
symptoms, vascular disease, and mild of Alzheimer disease. Arch. Neurol. 60, that are increased in Alzheimer’s disease.
cognitive impairment: findings from the 203–208 (2003). J. Biol. Chem. 279, 14673–14678 (2004).
Cardiovascular Health Study. Arch. Gen. 96. Huang, T. L. et al. Benefits of fatty fish on 115. Butterfield, D. A., Castegna, A., Drake, J.,
Psychiatry 63, 273–279 (2006). dementia risk are stronger for those without Scapagnini, G. & Calabrese, V. Vitamin E and
78. Becker, J. T. et al. Depressed mood is not a risk APOE ε4. Neurology 65, 1409–1414 (2005). neurodegenerative disorders associated with
factor for incident dementia in a community- 97. Kalmijn, S. et al. Dietary fat intake and the risk oxidative stress. Nutr. Neurosci. 5, 229–239
based cohort. Am. J. Geriatr. Psychiatry 17, of incident dementia in the Rotterdam Study. (2002).
653–663 (2009). Ann. Neurol. 42, 776–782 (1997). 116. Pitchumoni, S. S. & Doraiswamy, P. M. Current
79. Panza, F. et al. Impact of depressive symptoms 98. Schaefer, E. J. et al. Plasma status of antioxidant therapy for Alzheimer’s
on the rate of progression to dementia in phosphatidylcholine docosahexaenoic acid Disease. J. Am. Geriatr. Soc. 46, 1566–1572
patients affected by mild cognitive impairment. content and risk of dementia and Alzheimer (1998).
The Italian Longitudinal Study on Aging. Int. disease: the Framingham Heart Study. Arch. 117. Weisburger, J. H. Vitamin C and prevention of
J. Geriatr. Psychiatry 23, 726–734 (2008). Neurol. 63, 1545–1550 (2006). nitrosamine formation. Lancet 2, 607 (1977).
80. Aleisa, A. M., Alzoubi, K. H., Gerges, N. Z. & 99. Roberts, R. O. et al. Polyunsaturated fatty acids 118. Pardo, B., Mena, M. A., Fahn, S. & Garcia de
Alkadhi, K. A. Chronic psychosocial stress- and reduced odds of MCI: the Mayo Clinic Yebenes, J. Ascorbic acid protects against
induced impairment of hippocampal LTP: Study of Aging. J. Alzheimers Dis. 21, 853–865. levodopa-induced neurotoxicity on a
possible role of BDNF. Neurobiol. Dis. 22, 100. Solfrizzi, V. et al. Dietary intake of unsaturated catecholamine-rich human neuroblastoma cell
453–462 (2006). fatty acids and age-related cognitive decline: line. Mov. Disord. 8, 278–284 (1993).
81. Mayeux, R. et al. Synergistic effects of a 8.5-year follow-up of the Italian Longitudinal 119. Voko, Z., Hollander, M., Hofman, A., Koudstaal,
traumatic head injury and apolipoprotein‑ Study on Aging. Neurobiol. Aging 27, P. J. & Breteler, M. M. Dietary antioxidants and
epsilon 4 in patients with Alzheimer’s disease. 1694–1704 (2006). the risk of ischemic stroke: the Rotterdam Study.
Neurology 45, 555–557 (1995). 101. Engelhart, M. J. et al. Diet and risk of Neurology 61, 1273–1275 (2003).
82. Rasmusson, D. X., Brandt, J., Martin, D. B. dementia: does fat matter?: The Rotterdam 120. Calder, P. C. Polyunsaturated fatty acids,
& Folstein, M. F. Head injury as a risk factor in Study. Neurology 59, 1915–1921 (2002). inflammation, and immunity. Lipids 36,
Alzheimer’s disease. Brain Inj. 9, 213–219 102. Scarmeas, N., Stern, Y., Tang, M. X., Mayeux, R. 1007–1024 (2001).
(1995). & Luchsinger, J. A. Mediterranean diet and risk 121. Yehuda, S., Rabinovitz, S., Carasso, R. L.
83. Schofield, P. W. et al. Alzheimer’s disease after for Alzheimer’s disease. Ann. Neurol. 59, & Mostofsky, D. I. The role of polyunsaturated
remote head injury: an incidence study. 912–921 (2006). fatty acids in restoring the aging neuronal
J. Neurol. Neurosurg. Psychiatry 62, 119–124 103. Scarmeas, N. et al. Mediterranean diet and membrane. Neurobiol. Aging 23, 843–853
(1997). mild cognitive impairment. Arch. Neurol. 66, (2002).
84. Fleminger, S., Oliver, D. L., Lovestone, S., 216–225 (2009). 122. Abbott, R. D. et al. Walking and dementia in
Rabe-Hesketh, S. & Giora, A. Head injury as a 104. Scarmeas, N. et al. Physical activity, diet, and physically capable elderly men. JAMA 292,
risk factor for Alzheimer’s disease: the risk of Alzheimer disease. JAMA 302, 627–637 1447–1453 (2004).
evidence 10 years on; a partial replication. (2009). 123. Fratiglioni, L., Paillard-Borg, S. & Winblad, B. An
J. Neurol. Neurosurg. Psychiatry 74, 857–862 105. Scarmeas, N., Stern, Y., Mayeux, R. active and socially integrated lifestyle in late
(2003). & Luchsinger, J. A. Mediterranean diet, life might protect against dementia. Lancet
85. Mortimer, J. A. et al. Head trauma as a risk Alzheimer disease, and vascular mediation. Neurol. 3, 343–353 (2004).
factor for Alzheimer’s disease: a collaborative Arch. Neurol. 63, 1709–1717 (2006). 124. Scarmeas, N., Levy, G., Tang, M. X., Manly, J.
re-analysis of case–control studies. EURODEM 106. Feart, C. et al. Adherence to a Mediterranean & Stern, Y. Influence of leisure activity on the
Risk Factors Research Group. Int. J. Epidemiol. diet, cognitive decline, and risk of dementia. incidence of Alzheimer’s disease. Neurology 57,
20 (Suppl. 2), S28–S35 (1991). JAMA 302, 638–648 (2009). 2236–2242 (2001).
86. Guo, Z. et al. Head injury and the risk of AD in 107. Anstey, K. J., Mack, H. A. & Cherbuin, N. 125. Verghese, J. et al. Leisure activities and the risk
the MIRAGE study. Neurology 54, 1316–1323 Alcohol consumption as a risk factor for of dementia in the elderly. N. Engl. J. Med. 348,
(2000). dementia and cognitive decline: meta-analysis 2508–2516 (2003).
87. Mehta, K. M. et al. Head trauma and risk of prospective studies. Am. J. Geriatr. 126. Rovio, S. et al. Leisure-time physical activity at
of dementia and Alzheimer’s disease: the Psychiatry 17, 542–555 (2009). midlife and the risk of dementia and
Rotterdam Study. Neurology 53, 1959–1962 108. Kang, J. H., Cook, N., Manson, J., Buring, J. E. Alzheimer’s disease. Lancet Neurol. 4,
(1999). & Grodstein, F. A randomized trial of vitamin E 705–711 (2005).
88. Plassman, B. L. et al. Documented head injury supplementation and cognitive function in 127. Churchill, J. D. et al. Exercise, experience and
in early adulthood and risk of Alzheimer’s women. Arch. Intern. Med. 166, 2462–2468 the aging brain. Neurobiol. Aging 23, 941–955
disease and other dementias. Neurology 55, (2006). (2002).
1158–1166 (2000). 109. Yaffe, K., Clemons, T. E., McBee, W. L. 128. Colcombe, S. & Kramer, A. F. Fitness effects on
89. Hartman, R. E. et al. Apolipoprotein E4 & Lindblad, A. S. Impact of antioxidants, zinc, the cognitive function of older adults: a meta-
influences amyloid deposition but not cell loss and copper on cognition in the elderly: analytic study. Psychol. Sci. 14, 125–130
after traumatic brain injury in a mouse model a randomized, controlled trial. Neurology 63, (2003).
of Alzheimer’s disease. J. Neurosci. 22, 1705–1707 (2004). 129. Dishman, R. K. et al. Neurobiology of exercise.
10083–10087 (2002). 110. Petersen, R. C. et al. Vitamin E and donepezil Obesity 14, 345–356 (2006).
90. Franz, G. et al. Amyloid β 1–42 and tau in for the treatment of mild cognitive impairment. 130. Emery, C. F., Schein, R. L., Hauck, E. R.
cerebrospinal fluid after severe traumatic brain N. Engl. J. Med. 352, 2379–2388 (2005). & MacIntyre, N. R. Psychological and cognitive
injury. Neurology 60, 1457–1461 (2003). 111. Sano, M. et al. A controlled trial of selegiline, outcomes of a randomized trial of exercise
91. Morris, M. C. et al. Dietary intake of antioxidant alpha-tocopherol, or both as treatment for among patients with chronic obstructive
nutrients and the risk of incident Alzheimer Alzheimer’s disease. The Alzheimer’s Disease pulmonary disease. Health Psychol. 17,
disease in a biracial community study. JAMA Cooperative Study. N. Engl. J. Med. 336, 232–240 (1998).
287, 3230–3237 (2002). 1216–1222 (1997). 131. Fabre, C., Chamari, K., Mucci, P., Masse-Biron, J.
92. Engelhart, M. J. et al. Dietary intake of 112. Chiu, C. C. et al. The effects of omega‑3 fatty & Prefaut, C. Improvement of cognitive function
antioxidants and risk of Alzheimer disease. acids monotherapy in Alzheimer’s disease and by mental and/or individualized aerobic training
JAMA 287, 3223–3229 (2002). mild cognitive impairment: a preliminary in healthy elderly subjects. Int.
93. Masaki, K. H. et al. Association of vitamin E randomized double-blind placebo-controlled J. Sports Med. 23, 415–421 (2002).
and C supplement use with cognitive function study. Prog. Neuropsychopharmacol. Biol. 132. Kramer, A. F., Erickson, K. I. & Colcombe, S. J.
and dementia in elderly men. Neurology 54, Psychiatry 32, 1538–1544 (2008). Exercise, cognition, and the aging brain. J. Appl.
1265–1272 (2000). 113. Freund-Levi, Y. et al. Effects of omega‑3 fatty Physiol. 101, 1237–1242 (2006).
94. Laurin, D., Masaki, K. H., Foley, D. J., acids on inflammatory markers in 133. Lautenschlager, N. T. et al. Effect of physical
White, L. R. & Launer, L. J. Midlife dietary cerebrospinal fluid and plasma in Alzheimer’s activity on cognitive function in older adults at

150  |  MARCH 2011  |  VOLUME 7  www.nature.com/nrneurol


© 2011 Macmillan Publishers Limited. All rights reserved
REVIEWS

risk for Alzheimer disease: a randomized trial. evidence for involvement of SorL1 and the 172. Hansson, O. et al. Prediction of Alzheimer’s
JAMA 300, 1027–1037 (2008). retromer complex. J. Neurosci. 30, disease using the CSF Aβ42/Aβ40 ratio in
134. Angevaren, M., Aufdemkampe, G., 13110–13115 (2010). patients with mild cognitive impairment.
Verhaar, H. J., Aleman, A. & Vanhees, L. 155. Beecham, G. W. et al. Genome-wide association Dement. Geriatr. Cogn. Disord. 23, 316–320
Physical activity and enhanced fitness to study implicates a chromosome 12 risk locus (2007).
improve cognitive function in older people for late-onset Alzheimer disease. Am. J. Hum. 173. Ewers, M. et al. Multicenter assessment of
without known cognitive impairment. Cochrane Genet. 84, 35–43 (2009). CSF-phosphorylated tau for the prediction of
Database of Systematic Reviews, Issue 3. Art. 156. Bertram, L., McQueen, M. B., Mullin, K., conversion of MCI. Neurology 69, 2205–2212
No.: CD005381. doi:10.1002/14651858. Blacker, D. & Tanzi, R. E. Systematic meta- (2007).
CD005381.pub3 (2008). analyses of Alzheimer disease genetic 174. Andersson, C. et al. Differential CSF biomarker
135. Carlson, M. C. et al. Midlife activity predicts risk association studies: the AlzGene database. Nat. levels in APOE‑ε4-positive and -negative
of dementia in older male twin pairs. Alzheimers Genet. 39, 17–23 (2007). patients with memory impairment. Dement.
Dement. 4, 324–331 (2008). 157. Carrasquillo, M. M. et al. Genetic variation in Geriatr. Cogn. Disord. 23, 87–95 (2007).
136. Fratiglioni, L. & Wang, H. X. Brain reserve PCDH11X is associated with susceptibility to 175. Hoglund, K. et al. Prediction of Alzheimer’s
hypothesis in dementia. J. Alzheimers Dis. 12, late-onset Alzheimer’s disease. Nat. Genet. 41, disease using a cerebrospinal fluid pattern of
11–22 (2007). 192–198 (2009). C‑terminally truncated β‑amyloid peptides.
137. Acevedo, A. & Loewenstein, D. A. 158. Reiman, E. M. et al. GAB2 alleles modify Neurodegener. Dis. 5, 268–276 (2008).
Nonpharmacological cognitive interventions in Alzheimer’s risk in APOE ε4 carriers. Neuron 54, 176. Fagan, A. M. et al. Inverse relation between
aging and dementia. J. Geriatr. Psychiatry Neurol. 713–720 (2007). in vivo amyloid imaging load and cerebrospinal
20, 239–249 (2007). 159. Potkin, S. G. et al. Hippocampal atrophy as a fluid Aβ42 in humans. Ann. Neurol. 59,
138. Ball, K. et al. Effects of cognitive training quantitative trait in a genome-wide association 512–519 (2006).
interventions with older adults: a randomized study identifying novel susceptibility genes for 177. Buerger, K. et al. CSF phosphorylated tau
controlled trial. JAMA 288, 2271–2281 (2002). Alzheimer’s disease. PLoS One 4, e6501 protein correlates with neocortical
139. Unverzagt, F. W. et al. Effect of memory (2009). neurofibrillary pathology in Alzheimer’s disease.
impairment on training outcomes in ACTIVE. 160. Harold, D. et al. Genome-wide association study Brain 129, 3035–3041 (2006).
J. Int. Neuropsychol. Soc. 13, 953–960 (2007). identifies variants at CLU and PICALM 178. Buerger, K. et al. No correlation between CSF
140. Alzheimer Disease Mutation Database. associated with Alzheimer’s disease. Nat. tau protein phosphorylated at threonine 181
Alzheimer Disease & Frontotemporal Dementia Genet. 41, 1088–1093 (2009). with neocortical neurofibrillary pathology in
Mutation Database [online], http://www.molgen. 161. Lambert, J. C. et al. Genome-wide association Alzheimer’s disease. Brain 130, e82 (2007).
vib-ua.be/ADMutations/ (2010). study identifies variants at CLU and CR1 179. Engelborghs, S. et al. No association of CSF
141. De Strooper, B. et al. Deficiency of presenilin‑1 associated with Alzheimer’s disease. Nat. biomarkers with APOEε4, plaque and tangle
inhibits the normal cleavage of amyloid precursor Genet. 41, 1094–1099 (2009). burden in definite Alzheimer’s disease. Brain
protein. Nature 391, 387–390 (1998). 162. Bertrand, P., Poirier, J., Oda, T., Finch, C. E. 130, 2320–2326 (2007).
142. Scheuner, D. et al. Secreted amyloid β‑protein & Pasinetti, G. M. Association of 180. Fukumoto, H. et al. Age but not diagnosis is the
similar to that in the senile plaques of apolipoprotein E genotype with brain levels of main predictor of plasma amyloid β‑protein
Alzheimer’s disease is increased in vivo by the apolipoprotein E and apolipoprotein J (clusterin) levels. Arch. Neurol. 60, 958–964 (2003).
presenilin 1 and 2 and APP mutations linked to in Alzheimer disease. Brain Res. Mol. Brain Res. 181. Shafaati, M., Solomon, A., Kivipelto, M.,
familial Alzheimer’s disease. Nat. Med. 2, 33, 174–178 (1995). Bjorkhem, I. & Leoni, V. Levels of ApoE in
864–870 (1996). 163. Wyss-Coray, T. et al. Prominent cerebrospinal fluid are correlated with Tau and
143. Green, R. C. et al. Risk of dementia among white neurodegeneration and increased plaque 24S-hydroxycholesterol in patients with
and African American relatives of patients with formation in complement-inhibited Alzheimer’s cognitive disorders. Neurosci. Lett. 425, 78–82
Alzheimer disease. JAMA 287, 329–336 (2002). mice. Proc. Natl Acad. Sci. USA 99, (2007).
144. Gatz, M. et al. Role of genes and environments 10837–10842 (2002). 182. Schmand, B., Huizenga, H. M. & van Gool, W. A.
for explaining Alzheimer disease. Arch. Gen. 164. Baig, S. et al. Distribution and expression of Meta-analysis of CSF and MRI biomarkers for
Psychiatry 63, 168–174 (2006). picalm in Alzheimer disease. J. Neuropathol. detecting preclinical Alzheimer’s disease.
145. Corder, E. H. et al. Gene dose of apolipoprotein E Exp. Neurol. 69, 1071–1077 (2010). Psychol. Med. 40, 135–145 (2010).
type 4 allele and the risk of Alzheimer’s disease 165. Seshadri, S. et al. Genome-wide analysis of 183. Kosaka, T. et al. The beta APP717 Alzheimer
in late onset families. Science 261, 921–923 genetic loci associated with Alzheimer disease. mutation increases the percentage of plasma
(1993). JAMA 303, 1832–1840 (2010). amyloid-beta protein ending at A beta42(43).
146. Kurz, A. et al. Apolipoprotein E type 4 allele and 166. Wigge, P. et al. Amphiphysin heterodimers: Neurology 48, 741–745 (1997).
Alzheimer’s disease: effect on age at onset and potential role in clathrin-mediated endocytosis. 184. Schupf, N. et al. Elevated plasma
relative risk in different age groups. J. Neurol. Mol. Biol. Cell 8, 2003–2015 (1997). amyloid β‑peptide 1–42 and onset of dementia
243, 452–456 (1996). 167. Kelly, B. L. & Ferreira, A. Beta-amyloid disrupted in adults with Down syndrome. Neurosci. Lett.
147. Poirier, J. et al. Apolipoprotein E polymorphism synaptic vesicle endocytosis in cultured 301, 199–203 (2001).
and Alzheimer’s disease. Lancet 342, 697–699 hippocampal neurons. Neuroscience 147, 185. Mayeux, R. et al. Plasma Aβ40 and Aβ42 and
(1993). 60–70 (2007). Alzheimer’s disease: relation to age, mortality,
148. Farlow, M. R. et al. Impact of APOE in mild 168. Yang, S. et al. Comparative proteomic analysis and risk. Neurology 61, 1185–1190 (2003).
cognitive impairment. Neurology 63, 1898–1901 of brains of naturally aging mice. Neuroscience 186. van Oijen, M., Hofman, A., Soares, H. D.,
(2004). 154, 1107–1120 (2008). Koudstaal, P. J. & Breteler, M. M. Plasma Aβ1–40
149. Myers, R. H. et al. Apolipoprotein E ε4 169. Drewes, G., Ebneth, A., Preuss, U., and Aβ1–42 and the risk of dementia: a
association with dementia in a population-based Mandelkow, E. M. & Mandelkow, E. MARK, a prospective case–cohort study. Lancet Neurol.
study: the Framingham study. Neurology 46, novel family of protein kinases that 5, 655–660 (2006).
673–677 (1996). phosphorylate microtubule-associated proteins 187. Lopez, O. L. et al. Plasma amyloid levels and
150. Daw, E. W. et al. The number of trait loci in late- and trigger microtubule disruption. Cell 89, the risk of AD in normal subjects in the
onset Alzheimer disease. Am. J. Hum. Genet. 66, 297–308 (1997). Cardiovascular Health Study. Neurology 70,
196–204 (2000). 170. Starcevic, M. & Dell’Angelica, E. C. 1664–1671 (2008).
151. Rogaeva, E. et al. The neuronal sortilin-related Identification of snapin and three novel 188. Lui, J. K. et al. Plasma amyloid‑β as a biomarker
receptor SORL1 is genetically associated with proteins (BLOS1, BLOS2, and BLOS3/reduced in Alzheimer’s disease: the AIBL study of aging.
Alzheimer disease. Nature Genet. 39, 168–177 pigmentation) as subunits of biogenesis of J. Alzheimers Dis. 20, 1233–1242 (2010).
(2007). lysosome-related organelles complex‑1 189. Schupf, N. et al. Peripheral Aβ subspecies as
152. Reitz, C. et al. Meta-analysis of the association (BLOC‑1). J. Biol. Chem. 279, 28393–28401 risk biomarkers of Alzheimer’s disease. Proc.
between variants in SORL1 and Alzheimer’s (2004). Natl Acad. Sci. USA 105, 14052–14057 (2008).
disease. Arch. Neurol. 68, 99–106 (2011). 171. Morris, D. W. et al. Dysbindin (DTNBP1) and the 190. Teipel, S. J. et al. Relation of corpus callosum
153. Reitz, C. et al. SORCS1 alters APP processing biogenesis of lysosome-related organelles and hippocampal size to age in nondemented
and variants may increase Alzheimer’s disease complex 1 (BLOC‑1): main and epistatic gene adults with Down’s syndrome. Am. J. Psychiatry
risk. Ann. Neurol. doi:10.1002/ana.22308. effects are potential contributors to 160, 1870–1878 (2003).
154. Lane, R. et al. Diabetes-associated SorCS1 schizophrenia susceptibility. Biol. Psychiatry 63, 191. Karas, G. et al. Precuneus atrophy in early-
regulates Alzheimer’s amyloid‑β metabolism: 24–31 (2008). onset Alzheimer’s disease: a morphometric

NATURE REVIEWS | NEUROLOGY VOLUME 7  |  MARCH 2011  |  151


© 2011 Macmillan Publishers Limited. All rights reserved
REVIEWS

structural MRI study. Neuroradiology 49, 198. Small, G. W. et al. Current and future uses of Acknowledgments
967–976 (2007). neuroimaging for cognitively impaired patients. This work was supported by grants from the NIH and
192. Krasuski, J. S. et al. Volumes of medial temporal Lancet Neurol. 7, 161–172 (2008). the National Institute on Aging (grants
lobe structures in patients with Alzheimer’s 199. Silverman, D. H. et al. Positron emission R37-AG15473 and P01-AG07232), as well as
disease and mild cognitive impairment (and in tomography in evaluation of dementia: regional funding from The Blanchette Hooker Rockefeller
healthy controls). Biol. Psychiatry 43, 60–68 brain metabolism and long-term outcome. JAMA Foundation and The Charles S. Robertson Gift from
(1998). 286, 2120–2127 (2001). the Banbury Fund (R. Mayeux), and a Paul B.
193. Mungas, D. et al. Longitudinal volumetric MRI 200. O’Brien, J. T. Role of imaging techniques in the Beeson Career Development Award (K23AG034550)
change and rate of cognitive decline. Neurology diagnosis of dementia. Br. J. Radiol. 80, to C. Reitz.
65, 565–571 (2005). S71–S77 (2007). Laurie Barclay, freelance writer and reviewer, is the
194. Apostolova, L. G. et al. Conversion of mild 201. Klunk, W. E. et al. Imaging brain amyloid in author of and is solely responsible for the content
cognitive impairment to Alzheimer disease Alzheimer’s disease with Pittsburgh of the learning objectives, questions and answers of
predicted by hippocampal atrophy maps. Arch. Compound‑B. Ann. Neurol. 55, 306–319 (2004). the MedscapeCME-accredited continuing medical
Neurol. 63, 693–699 (2006). 202. Engler, H. et al. Two-year follow-up of amyloid education activity associated with this article.
195. Likeman, M. et al. Visual assessment of atrophy deposition in patients with Alzheimer’s disease.
on magnetic resonance imaging in the diagnosis Brain 129, 2856–2866 (2006).
Author contributions
of pathologically confirmed young-onset 203. Frisoni, G. B. et al. In vivo mapping of amyloid
C. Reitz, C. Brayne and R. Mayeux researched the
dementias. Arch. Neurol. 62, 1410–1415 (2005). toxicity in Alzheimer disease. Neurology 72,
data for the article, provided substantial contributions
196. Chetelat, G. et al. Using voxel-based morphometry 1504–1511 (2009).
to discussions of the content, and contributed equally
to map the structural changes associated with 204. Tolboom, N. et al. Differential association of
to writing the article and to review and/or editing of
rapid conversion in MCI: a longitudinal MRI study. [11C]PIB and [18F]FDDNP binding with cognitive
the manuscript before submission.
Neuroimage 27, 934–946 (2005). impairment. Neurology 73, 2079–2085 (2009).
197. Rombouts, S. A. et al. Functional MR imaging in 205. Katzman, R. Editorial: the prevalence and
Alzheimer’s disease during memory encoding. malignancy of Alzheimer disease. A major killer. Supplementary information is linked to the online
AJNR Am. J. Neuroradiol. 21, 1869–1875 (2000). Arch. Neurol. 33, 217–218 (1976). version of the paper at www.nature.com/nrneurol

152  |  MARCH 2011  |  VOLUME 7  www.nature.com/nrneurol


© 2011 Macmillan Publishers Limited. All rights reserved

You might also like