Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2009 Aug 1.
Published in final edited form as: Lancet Neurol. 2008 Aug;7(8):704–714. doi: 10.1016/S1474-4422(08)70162-5

Biomarkers for cognitive impairment and dementia in elderly people

Joshua A Sonnen 1, Kathleen S Montine 1, Joseph F Quinn 2, Jeffrey A Kaye 2, John CS Breitner 3, Thomas J Montine 1,2
PMCID: PMC2710375  NIHMSID: NIHMS115055  PMID: 18635019

Abstract

The threat of a looming pandemic of dementia in elderly people highlights the compelling need for the development and validation of biomarkers that can be used to identify pre-clinical and prodromal stages of disease in addition to fully symptomatic dementia. Although predictive risk factors and correlative neuroimaging measures will have important roles in these efforts, this Review describes recent progress in the discovery, validation, and standardization of molecular biomarkers – small molecules and macromolecules whose concentration in brain or biological fluids can aid in diagnosis at different stages of the more common dementing diseases and in the assessment of disease progression and response to therapeutics. An approach that efficiently combines independent information from risk factor assessment, neuroimaging measures, and biomarkers may soon guide clinicians in the early diagnosis and management of cognitive impairment in elderly people.

INTRODUCTION

Cognitive impairment and dementia in elderly people represent a burgeoning public health problem that already causes untold suffering and threatens to overwhelm health care delivery systems in the coming decades. One response to the looming pandemic of dementing illness is the development of biomarkers that can aid in diagnosis, prognosis, selection for clinical trials, and objective assessment of therapeutic response. Here we review the current level of knowledge for biomarkers of the common dementing illnesses, paying particular attention to their applicability at different stages of disease progression and to the quality of the current evidence.

PATHOGENESIS OF COGNITIVE IMPAIRMENT

Progressive decline in cognitive function among elderly people results mostly from Alzheimer’s disease (AD) and vascular cognitive impairment (VCI),1 and less frequently from Lewy body disease (LBD)2 or frontotemporal dementia (FTD);3 these forms of dementia combine to varying extents in individual patients.49 The cellular and molecular pathogenesis of the common dementing illnesses is known only in part. For AD, LBD and FTD, complex molecular cascades seem to derive from the accumulation of abnormal forms of proteins or peptides, which suggests a proposed class of protein-misfolding diseases or so called proteinopathies.10,11 Key among these proteins are amyloid β (Aβ) peptides and tau in AD, α-synuclein in LBD, and tau or TAR DNA-binding protein (TARDBP; alias TDP-43) in forms of FTD. Important pathogenic components in these diseases include innate immune activation, excitotoxicity, mitochondrial dysfunction, and oxidative damage, all of which may lead to regional loss of synapses and, ultimately, to loss of neurons.

STAGES OF DISEASE PROGRESSION

More than 30 years ago, Robert Katzman12 proposed a chronic disease model for AD that suggests that the disease progresses from a preclinical or latent stage with some structural or molecular damage but no functional or behavioral changes, through a prodromal stage with greater damage and mild functional or behavioral changes, to fully expressed clinical syndrome of dementia, typically provoked by substantial and irreversible damage. This perspective is now supported strongly by clinical, neuroimaging and autopsy data.49,1319 Although they are less complete, emerging data similarly indicate that a chronic disease model may apply to VCI, LBD, and FTD.1,4,5,18,19 Although these theoretical disease stages are evident in longitudinal studies, they are less apparent in cross-sectional designs. Table 1 shows some of the approaches commonly used to categorize patients into different stages of dementing illnesses in cross-sectional biomarker studies. One approach uses measures such as the clinical dementia rating scale.24 Another is ascertainment of discrete categories that mainly compromise individuals with disease at different stages. For example, amnestic mild cognitive impairment (MCI) seems to constitute a group that largely consists of patients with prodromal AD,20 whereas non-amnestic MCI is a proposed, but less accurate, indicator of prodromal VCI.25 Finally, there are several subtypes of the dementia syndrome with established clinical criteria that have varying levels of sensitivity and specificity for the underlying disease process. For example, vascular dementia (VaD) is a clinically defined category that mainly inlcudes patients with dementia-stage illness from VCI. Similarly, as applied in several biomarker studies, dementia with Lewy bodies (DLB) is evolving as a separate, clinically-defined category for patients with dementia-stage illness from that of LBD.

Table 1.

Some commonly used approaches to categorize groups of individuals at different stages of dementing illnesses for biomarker studies

Pre-clinical Prodrome Dementia
Alzheimer’s disease PSEN1 mutations, trisomy 21, APOEε4, family history, age Amnestic MCI20 Probable AD21
Vascular cognitive impairment Risk factors for vascular disease, neuroimaging, clinical and family history, age Non-amnestic MCI20 Vascular dementia22
Lewy body disease nr nr Dementia with Lewy bodies23
Frontotemporal dementia nr nr ¨ Frontotemporal

AD=Alzheimer’s disease. MCI=Mild cognitive impairment. nr =not reported.

RISK FACTORS, BIOMARKERS AND DISEASE SURROGATES

Risk factors are identifiable events or conditions associated with an increased probability of disease. For example, advanced age is a risk factor for AD. Genes can also be risk factors, as is shown dramatically by heritability estimates of 0.6 or higher from recent twin studies of AD.13 Inheritance of the ε4 allele of the apolipoprotein E gene (APOE) is associated with earlier age of onset of AD than is inheritance of the other two common alleles of this gene.26 Accordingly, age and inheritance of the ε4 allelehave been used in biomarker studies to increase of including individuals with pre-clinical AD (see Table 1 and Table 2). We define biomarkers as molecules whose concentration in brain or biological fluids aids diagnosis, assessment of disease progression, or response to therapeutics. Thus, biomarker research in geriatric dementia focuses on quantification of specific molecules using techniques that range from PET ligand binding to gene expression microarrays. Under this definition, structural and functional neuroimaging do not qualify as biomarkers, even though they contribute important information to the assessment of cognitively impaired patients. Disease surrogate is a concept from clinical trials that describes a trait or measure used as a substitute for the clinical endpoint of interest, often because the latter is difficult or unethical to obtain.

Table 2.

Studies of biomarkers in Alzheimer’s disease

n Main Outcome Stage
CSF T-tau or Aβ42
Peskind and co-workers16 184 ↓ Aβ42 in patients aged <50 years, most prominently in APOE ε4 carriers Preclinical
Kauwe and co-workers17 191 ↓ Aβ42 or Aβ42: Aβ40 was highly correlated with age, CDR of 0–1, and APOE ε4; PSEN1 mutation identified in a patient with greatly increased Aβ40 and Aβ42 concentrations Preclinical, prodrome, dementia
Sunderland and co-workers27 292 ↓ Aβ42, but no change in tau in APOE ε4 carriers and ↓ Aβ42 and ↑ tau in AD dementia Preclinical, dementia
Fagan and co-workers28 139 ↓ Aβ42 and ↑ tau in patients with CDR ≥0.5 Prodrome, dementia
Lewczuk and co-workers29 223 ↓ Aβ42 and ↑ tau in amnestic MCI and AD dementia compared with non-amnestic MCI; ↑ tau in AD dementia compared with amnestic MCI Prodrome, dementia
Sunderland and co-workers30 203 ↓ Aβ42 and ↑ tau in AD Dementia
Andreasen and co-workers31 241 ↓ Aβ42 and ↑ tau in AD dementia; ↑ tau in AD dementia compared with VaD, MCI, DLB, and controls Prodrome, dementia
Andreasen and co-workers32 111 ↑ tau in AD and some VaD Dementia
Galasko and co-workers33 216 ↓ Aβ42 and ↑ tau in AD; APOE ε4 gene-dosage effect on Aβ42 concentrations Dementia
Bouwman and co-workers34 375 ↓Aβ42 and ↑ T-tau in AD dementia and ↓ Aβ42 in older controls Dementia
Hansson and co-workers35 * 137 ↓Aβ42 and Aβ42:Aβ40 in MCI that progressed to AD over 4–6 years Prodrome
Ewers and co-workers36 * ↑ T-tau:Aβ42 predicted progression to MCI or AD dementia over up to 40 months Preclinical
CSF tau-P181
Fagan and co-workers28 139 ↑ in patients with CDR ≥0.5 Prodrome, dementia
Lewczuk and co-workers29 223 ↑ in amnestic MCI and AD dementia compared with non-amnestic MCI Prodrome, dementia
Bouwman and co-workers34 375 ↑ in AD compared with controls and in older controls compared with younger controls Dementia
CSF tau-P231
Glodzik-Sobanska and co-workers37 78 ↑ in APOE ε4 carriers Preclinical
Buerger and co-workers38 131 ↑ in MCI APOE ε4 carriers Prodrome
Buerger and co-workers39 192 ↑ in AD compared with controls, FTD, VaD, and DLB Dementia
Ewers and co-workers36 * 145 ↑ in MCI that converted to AD over 14–28 mos Prodrome
CSF F2-IsoPs
Glodzik-Sobanska and co-workers37 78 APOE ε4 carriers Preclinical
Pratico and co-workers40 63 ↑ in MCI and AD Prodrome, dementia
de Leon and co-workers41 16 ↑ in MCI Prodrome
Montine and co-workers4245 Pratico and co-workers46 14–77 ↑ in AD and HD but not in ALS or MSA Dementia
Quinn and co-workers15 * de Leon and co-workers47 * 40,16 ↑ over 1 year in patients with CDR 0.5–1 Dementia
CSF MAP
Hu and co-workers48 94 6-member proteomics-discovered MAP effective in identifying patients with CDR 0.5–1 Prodrome, dementia
Zhang and co-workers49 183 8-member proteomics-discovered MAP moderately effective in identifying AD, strongly effective for PD Dementia
Simonsen and co-workers50 137 15-member proteomics-discovered group validated to differentiate AD from controls Dementia
11C-PiB-PET
Mintun and co-workers51 51 ↑ binding in AD dementia and a subset of asymptomatic individuals Preclinical, dementia
Pike and co-workers52 96 ↑ binding in 97% of AD dementia, 61% of MCI, and 22% of controls, and was correlated with objective deficits in episodic memory Preclinical, prodrome, dementia
Fagan and co-workers53 24 ↑ binding was correlated with ↓ CSF Aβ42 in individuals with CDR of 0–2 Preclinical, prodrome, dementia
Rabinovici and co-workers54 27 ↑ binding in AD and some FTD cases Dementia
Archer and co-workers55 9 Binding correlated with brain atrophy in AD Dementia
Edison and co-workers56 33 ↑ binding in 17 of 19 AD patients and correlated with reduced glucose uptake by FDG-PET in some regions Dementia
Klunk and co-workers57 25 ↑ binding in AD Dementia
42:Aβ40 plasma ratio
van Oijen and co-workers58 1756 ↓ Aβ42:Aβ40 associated with increased risk for AD and VaD Dementia
Graff-Radford and co-workers59 * 563 ↓ ratio at baseline associated with increased risk for conversion to MCI or AD dementia at 3–7 years Preclinical
Sundelof and co-workers60 * 1725 ↓ Aβ40 predicted conversion to AD dementia in 77-year-old men over up to 7.9 years Preclinical
Lopez and co-workers61 * 274 ↓ Aβ42, ↑ Aβ40, and ↓ Aβ42: Aβ40 predicted conversion from control or MCI to AD, but this association was lost in models that corrected for vascular disease risk Preclinical, prodrome
*

Longitudinal studies (all others are cross-sectional studies).

↑=biomarker increased. ↓ =biomarker decreased. Aβ=amyloid β. AD=Alzheimer’s disease. ALS=amyotrophic lateral sclerosis. CDR=clinical dementia rating (score). DLB=dementia with Lewy bodies. FDG= fluorodeoxyglucose. FTD=frontotemporal dementia. HD=Huntington’s disease. IsoPs=isoprostanes. MAP=multianalyte profile. MCI=mild cognitive impairment. MSA=multisystem atrophy. PD=Parkinson’s disease. PiB=Pittsburgh compound B. tau-P=phosphorylated tau. T-tau=total tau. VaD=vascular dementia.

Before reviewing the existing biomarkers for geriatric dementia, we must discuss the nature and quality of the data. We suggest a hierarchy of biomarker development, progressing from discovery of initial associations, to confirmation in a larger or more complex independent sample, to validation with a standardized quantitative assay in other laboratories, and finally to widespread clinical application of that assay. Whereas initial studies might appropriately use bespoke assays to discover initial associations, validation requires uniform assays with authentic standards, quality assurance, and known performance characteristics. To this evidence, then, must be added knowledge of potential effects of other common illnesses and commonly used drugs or supplements,14,15 age,16,17 sex,62 diet, level of physical activity, and circadian variation.63 With standardization and rigorous quality control, markers may finally be adopted for widespread application in clinical laboratories.

BIOMARKERS FOR ALZHEIMER’S DISEASE

Most data for biomarkers of geriatric dementia are for AD, usually diagnosed as “probable AD” by expert physicians and neuropsychologists in tertiary medical centers.21 Many of these centers have active research programs that include autopsy and neuropathologic classification of dementing disorders. Therefore, the performance characteristics for expert clinical diagnosis of AD dementia are known. In general, the sensitivity and positive predictive value (vs neuropathological confirmation) for a clinical diagnosis of “probable AD” are about 90 to 95%, but specificity is lower (about 50 to 60% ) because commonly co-morbid VCI and LDB are often difficult to discern clinically. This limitation of specificity presents an important problem for the interpretation of biomarker studies that have relied on expert clinical diagnosis: the probable AD group is almost always a mixture of patients with dementia that derives from AD alone and a substantial group that has additional dementing illnesses, most often VCI or LBD. One method of addressing this problem is the limitation of biomarker studies to individuals who subsequently undergo autopsy for “gold standard” classification of dementing illnesses.6466 Besides greatly limiting the number of cases, this approach assumes that each neuropathologic process present at the time of death was also present earlier when the biomarker was quantified. Considering these limitations, we have summarized current knowledge about biomarkers for AD identified in (A) cross-sectional and (B) longitudinal studies (table 2). Although direct comparison of the results from different laboratories is difficult, the similarity of outcomes across studies conveys the extent of validation at each stage of disease. .

CSF Biomarkers

Investigations of biomarkers for AD are dominated by cross-sectional studies for CSF Aβ42 and tau species at each stage of disease. These markers have achieved widespread application in AD dementia and are currently being validated in prodromal AD or MCI. However, in pre-clinical AD, research is still at the stage of discovery.

CSF Aβ42 concentrations are decreased by about 50% in patients with AD dementia or MCI, compared with (typically age-matched) controls.67 This decrease has been associated with enhanced deposition of Aβ42 in the brain.28 By contrast, CSF total tau (T-tau) is increased on average by two to three times in AD dementia and MCI, whereas some phosphorylated tau (tau-P) species, (e.g., tau-P231 or tau-P181) can be increased by one or two orders of magnitude when compared to control levels.29,31,34,38,39 Since reduced CSF Aβ42 and increased CSF tau are both characteristic changes for prodrome AD and dementia, investigators often combine the results of both assays.33,67 Indeed, the CSF tau/Aβ42 ratio seems to distinguish patients with very slight cognitive impairment from controls.28 Unfortunately, none of these changes in CSF protein concentration is specific to AD,68 and the biomarker levels are highly variable across individuals such that nearly all studies show substantial overlap in CSF values among controls and patient groups.3032,34,38,39,64,69 Moreover, especially for CSF tau species, the relationship between the concentration of these CSF proteins and the actual burden of disease is unclear.70 The extent to which these limitations indicate misclassifications of patients versus true discrepancies cannot be resolved at this point and can be addressed only partially with autopsy classification, as noted above.

Because there are no characteristic symptoms in pre-clinical AD, cross-sectional investigations at this stage are challenging. Most investigators have used inherited mutations or risk factors to identify healthy older adults who are nevertheless likely to develop subsequent cognitive impairment. Some,27 but not all,37 groups have observed that CSF Aβ42 is decreased to dementia-like levels in cognitively normal elderly with an APOE ε4 allele. Another group observed an age-related (cross-sectional) decline in CSF Aβ42 in cognitively normal individuals beginning in the sixth decade of life, with a notable exaggeration of this decline in those with APOE ε4.16 These results suggest that decreased CSF Aβ42 may be a biomarker of pre-clinical AD. By contrast, others recently identified a cognitively normal volunteer with a family history of AD who had exceptionally high CSF Aβ42 concentration and was shown to have inherited a disease-causing mutation in PSEN1 (presenilin 1).17 These results indicate that pre-clinical AD itself might have different stages (early increased CSF Aβ42 with little parenchymal deposition followed by decreased CSF Aβ42 secondary to parenchymal deposition). Alternatively, the pattern of changes in CSF biomarkers might differ among those with APOE ε4 versus a disease-causing mutation. Some,37,38 but not all,27 investigators have observed increased CSF tau or tau-P231 concentrations in cognitively normal older adults who inherited an APOE ε4 allele, suggesting that increased CSF tau and tau-P231 could be biomarkers of at least one stage or form of pre-clinical AD.

In longitudinal analyses of CSF Aβ42 and tau, many investigators have shown that decreased CSF Aβ42 and increased CSF T-tau or phosphorylated tau species predict subsequent conversion from MCI to AD dementia over follow up periods extending to 6 years,35,36,7176 while one group has observed similar predictive power for the relative pattern of C-terminal truncated Aβ peptides.77 Others have reported that patients with MCI or AD dementia show cross-sectional alterations in CSF Aβ42, T-tau, and tau-P181 concentrations that far exceed interval changes over a mean follow up period of 21 months.78 These findings suggest that these measures are not particularily sensitive markers of disease progression at the prodromal or dementia stages of AD. Initial studies of empirically defined CSF tau/Aβ42 ratio, however,suggest that these measures could prove helpful in identifying pre-clinical AD.28,79

Another extensively studied CSF biomarker is concentration of F2-isoprostanes. Unlike disease-oriented (if not disease-specific) biomarkers like Aβ42 and tau, F2-IsoPs are mechanism-specific – that is, they are quantitative markers of free radical damage to lipids in vivo. In cross-sectional studies, increased CSF concentrations of this biomarker have been validated in AD dementia,15,40,4246 and similar findings have been confirmed in MCI.40,41 Initial associations have also been made in individuals with an APOE ε4 allele.37 Thus, CSF F2-isoprostanes could prove to be useful biomarkers at all stages of AD pathogenesis, a possibility that is further supported by results from brain samples from patients who died with AD at prodromal or dementia stages that show increased F2-isoprostanes or related molecules.8082 Longitudinal studies have also found a 1-year interval increase in CSF F2-isoprostanes concentrations in patients with early AD dementia that might be suppressed by anti-oxidant vitamin supplements.15,58

Several groups are now using relatively unbiased discovery approaches to identify CSF biomarkers for neurodegenerative diseases. One method used by several laboratories is proteomics of CSF in cross-sectional studies.66,8386 Several partially overlapping ensembles of CSF proteins have been identified from patients with AD dementia. However, the number of patient samples is typically small and only twice have diagnoses been confirmed by autopsy.66,84 A few groups have confirmed their initial findings.4850,87 One group used surface-enhanced laser desorption ionization to identify a panel of 17 potential biomarkers that distinguished a group of patients with MCI who progressed to dementia over 4 to 6 years.87 Some molecular species in this panel were subsequently confirmed in analyses of separate samples.50 Similarly, Hu, and colleagues48 have assembled a proteomics-discovered six member multi-analyte profile (MAP) of CSF proteins that partially distinguished patients with very mild AD from controls. Another group has developed a CSF proteomics-discovered eight member MAP and confirmed its use in a relatively large number of individuals (n=202) from multiple centers with different neurodegenerative diseases.49

In summary, CSF biomarker research in AD remains dominated by cross-sectional studies of Aβ peptides and tau species, which now have extensive validation in both dementia and prodromal stages. CSF F2-IsoPs are validated in AD dementia and in MCI. Longitudinal studies and application of CSF proteomic approaches are still in discovery or confirmation phases.

Plasma or urine biomarkers

Although there is clear rationale for pursuing AD biomarkers by PET probes or in CSF,69,88 there is no question that a biomarker derived from a peripheral fluid would be vastly more convenient and would probably aid the execution of large clinical trials. Unfortunately, plasma or urine biomarkers remain far less developed than those in CSF. Repeatedly published observations of increased plasma or urine concentrations of F2-isoprostanes in patients with MCI or AD dementia could not be confirmed by others89, and the original findings were not replicated by the same investigators using a different sample set.90

The data on concentrations of plasma Aβ species are complicated. Increased plasma Aβ42 and sometimes Aβ40 have been reported in rare autosomal dominant early-onset forms of AD and in patients with Down’s syndrome.9193 Other investigators have repeatedly observed increased plasma Aβ42 but not Aβ40 concentrations in patients with late-onset AD dementia;9496 however, the usefulness of increased plasma Aβ42 as a biomarker for this much more common form of AD is not clear.97 Indeed, a recent cohort analysis from the Rotterdam Study concluded that increased plasma concentrations of Aβ40, especially when combined with reduced plasma levels of Aβ42, are associated with a risk of dementia from AD or VaD that is increased by up to ten times.58 As with CSF, there is substantial overlap in plasma Aβ42 concentrations among controls and patients with AD or other forms of dementia in cross sectional studies. One group did longitudinal analysis of cognitively normal older adults and concluded that reduced plasma Aβ42:Aβ40 ratio is associated with an increased risk of subsequently developing MCI or AD dementia.59 However, a recent study of plasma Aβ peptides in 1725 men in their seventies concluded that decreased plasma Aβ40 predicts incident AD.60 A recently published 4.5 year longitudinal study of plasma Aβ peptide concentrations in a subset of participants in the Cardiovascular Health Study Cognition Study who were normal (n=242) or had MCI (n=42) concluded that, after controlling for age and other illnesses including cerebrovascular disease, plasma Aβ40, Aβ42, and their ratio did not seem to be useful biomarkers for AD.61 A final level of complexity is raised by a recent study which showed 11C-Pittsburgh compound B (PiB)-PET imaging correlates well with CSF Aβ42 concentration, but that there is no discernible trend toward correlation with plasma concentrations of Aβ species.53 Obviously, a deeper understanding is needed of the mechanistic connections among the brain, CSF, and plasma Aβ species in health and disease.

Several other plasma biomarkers for AD have been proposed, mostly related to immune activation or oxidized low-density lipoproteins. The specificity of these peripheral markers of inflammation has not been rigorously studied and they remain mostly in the discovery phase of development.98102 Increases in plasma α1-antichymotrypsin has been validated in AD dementia with varying degrees of correlation to disease severity,103,104 but this finding was not supported by a third study.105 A recent report identified 18 signaling peptides in plasma (out of 120 candidates), with concentrations that identified patients with AD dementia in a cross-sectional analysis, and progression from prodrome to dementia stage AD.106 This is a potentially very important study; however, as with previously proposed plasma biomarkers of AD, independent validation of this panel's performance will be key to its future application.

In summary, biomarkers of AD in plasma, serum, or urine are highly desirable and thus are an area of active investigation that is still largely in the discovery phase. Several initially promising discoveries have not withstood attempts at validation, whereas others await validation.

PET for cerebral amyloid

Several amyloid-binding compounds can be synthesized with unstable isotopes compatible with radiotracer imaging. The most widely reported technique has used 11C-labeled Pittsburgh Compound-B (2-[4L’-(methyl-amino)phenyl]-6-hydrobenzothiazole; 11C-PiB) for PET imaging of amyloid binding.57 The 11C-PiB ligand is distributed rapidly to all grey-matter regions in the brain and is selectively retained in regions where it binds. Typically, data from cerebral regions are normalized to 11C-PiB retention in cerebellum. Average and regional 11C-PiB-PET signals are inversely correlated with MRI volumetric measurements, but do not co-register well with 18F-FDG-PET in the frontal lobe of patients with AD dementia.55,56 Although some have referred to PiB as a probe for amyloid plaque binding,53 the ligand binds to amyloid deposits not only in the parenchyma, but also in cerebral blood vessels107 that apparently can be the major source of signal.107,108 The mean 11C-PiB-PET signal intensity from multiple cerebral cortical regions is inversely correlated with CSF Aβ42 concentrations, but not with CSF Aβ40, CSF T-tau, CSF tau-P181, plasma Aβ42, or plasma Aβ40 in individuals with no cognitive impairment to moderate AD dementia.53 This important demonstration of the relevance of CSF Aβ42 concentrations to the burden of disease in the brain is lacking for other CSF or plasma biomarkers. Along with other reports, these data support the hypothesis that decreased CSF Aβ42 is a consequence of increased cerebral deposition that may begin in pre-clinical AD17,51,53 where it may be associated with diminished (if still normal) episodic memory.52

Sensitivity of 11C-PiB-PET against a standard of expert clinical diagnosis of AD dementia is high (about 90%)51,56 but autopsy investigations will be needed to assess this finding further. Specificity of the 11C-PiB-PET signal for AD and congophilic amyloid angiopathy (vs other changes) is currently under investigation. PiB does not seem to bind to cortical Lewy bodies,109 and PET imaging of 11C-PiB retention in a patient with an autosomal dominant form of prion disease was similar to that in cognitively normal controls.110 However, PET imaging with another amyloid radiotracer, FDDNP(2-(1-(6-[2-[18F]fluoroethyl)(methyl)amino]-2-naphthyl)ethylidene) malononitrile), showed values between those from controls and AD in a sibling who had the same mutation. The latter observation suggests that different amyloid binding radiotracers may vary in their evaluation of dementing illnesses.110 Indeed, results with 18F-BAY94-9172, a novel Aβ ligand and PET tracer, were recently reported in 15 controls, 15 patients with AD, and 5 patients with FTD. Blinded reading of PET images correctly distinguished all FTD patients from those with AD and had 100% sensitivity and 90% specificity for the clinical diagnosis of AD or control.111

BIOMARKERS FOR VASCULAR COGNITIVE IMPAIRMENT

The same difficulties that surround expert clinical diagnosis of AD also exist also for VCI, perhaps even more so.1 As we noted above, this limitation in clinical diagnosis can serve as a motivation for biomarker development, but it also imposes a considerable challenge to such work because reliable diagnosis is more dependent on neuropathologic assessment with all its limitations.

VCI spans from local territorial infarcts, for which structural MRI provides unsurpassed insight into lesion size and progression, to widespread small vessel disease, which isnot as easily assessed or distinguished from the consequences of AD by clinical examination or by current standard neuroimaging techniques.112 Thus, biomarkers could contribute to the assessment of damage from small vessel disease, whereas neuroimaging is likely toretain its central role in the assessment of ischemic injury from large and medium caliber vessels.

In 2006 the US National Institute for Neurological Disorders and Stroke and the Canadian Stroke Network made expert recommendations for biomarkers in VCI.22 Increased CSF concentrations of T-tau or phosphorylated tau species, sometimes combined with normal Aβ42 concentrations, increased neurofilament protein, and increased CSF:serum albumin ratio all have been reported in patients with dementia in cross-sectional studies.31,113117 Unfortunately, all such findings are limited by the broad overlap with values from patients whose clinical appearance suggests AD without VCI (Table 3). CSF sulfatide concentrations seem to be raised in VaD but have been shown only once to differ significantly different from those found in AD dementia.119,120 CSF matrix metalloproteinase is raised in VaD but not AD dementia or controls.121 One longitudinal study found that normal CSF tau levels combined with MRI evidence of presumed white matter ischemic injury could be used to identify patients with MCI who are less likely to progress to dementia.118

Table 3.

Studies of Biomarkers in vascular cognitive impairment

n Main outcome Stage
CSF t-tau and Aβ42
Andreasen and co-workers31 241 ↑ tau, but normal Aβ42 in VaD Dementia
Stefani and co-workers113 110 ↓ Aβ42 in AD discriminates from VaD Dementia Dementia
Maruyama and co-workers118 * 57 Stable MCI was characterized by normal T-tau concentrations and high-grade periventricular white-matter lesions Prodrome
CSF tau-P181
Stefani and co-workers113 110 ↑ in AD dementia discriminates from VaD Dementia
CSF sulphatide
Tullberg and co-workers119 90 ↑ in subcortical arteriosclerotic encephalopathy Dementia
Fredman and co-workers120 83 ↑ in VaD Dementia
CSF neurofilament protein
Sjogren and co-workers117 51 ↑ in VaD Dementia
Wallin and co-workers116 43 ↑ in patients with dementia with several vascular risk factors or white-matter imaging abnormalities Dementia
CSF MMP9
Adair and co-workers121 53 ↑ in VaD compared with controls and AD dementia Dementia
CSF:serum albumin ratio
Skoog and co-workers115 65 ↑ in VaD and AD compared with controls; ↑ in women (n=3) who progressed to dementia over 3-year follow-up Preclinical, dementia
Wallin and co-workers116 43 ↑ in VaD Dementia
Plasma Aβ40
van Oijen and co-workers58 1756 ↑ Aβ40 (and ↓ Aβ42) associated with increased risk for AD dementia and VaD Dementia
Bibl and co-workers122 72 ↑ in VCI Dementia
38:Aβ40 plasma ratio
Bibl and co-workers122 72 ↓ in VCI Dementia
42:Aβ40 plasma ratio
Sundelof and co-workers60 * 1725 1 SD ↑ in Aβ42:Aβ40 ratio associated with increased risk for VCI in 70-year-old men Preclinical
*

Longitudinal studies (all others are cross-sectional studies).

↑ =biomarker increased. ↓ =biomarker decreased. Aβ=amyloid β. AD=Alzheimer’s disease. MCI=mild cognitive

Several plasma-based or serum-based biomarkers for VCI have also been proposed. One recent longitudinal study discovered that increased plasma fibrinogen, but not C-reactive protein was associated with an increased risk for VaD and AD dementia.100 Another study found that the combination of high C-reactive protein and high interleukin-6 concentrations increased the risk of incident VaD, but not AD dementia.100,123 As in AD, the evidence for plasma Aβ peptide biomarkers for VCI is not entirely consistent. One recent study found that a 1 SD increase in serum Aβ42:Aβ40 ratio increased risk for VaD in 70-year-old men but not 77-year-old men.60 Another group found plasma Aβ40 to be increased and the Aβ38:Aβ40 ratio decreased in VCI.122 In summary, biomarker development in VCI is less mature than in AD and has not yet identified candidates that are ready for standardization and widespread application.

BIOMARKERS FOR LEWY BODY DISEASE

LBD is characterized by regional intraneuronal accumulation of Lewy bodies composed of several proteins that prominently include α-synuclein. Accumulation of Lewy bodies in neocortical regions of the brain is a strong correlate of dementia.4 However, clinical criteria that can reliably distinguish DLB from AD dementia are still under development. 23 As a result, biomarker studies of patients without autopsy confirmation of diagnosis are likely to rely on groups that mainly compromise patients with AD or DLB, but with an unknown degree of diagnostic cross-contamination.114

We are unaware of biomarker studies that focus on prodromal or pre-clinical DLB. One large (> 30 individuals per group) multicenter cross-sectional study of CSF biomarkers for DLB versus AD used commercially available kits to measure CSF Aβ42, T-tau, and tau-P181. This study concluded that CSF tau-P181 is the most useful marker for distinguishing DLB from AD dementia, yielding an overall accuracy of 80%.124 These findings confirmed a smaller previous study.114 Another group has reported that a relative abundance of Aβ37 and Aβ42 in CSF discriminates between relatively small samples of patients with DLB and AD dementia.125 Efforts directed at biomarkers for another form of LBD, Parkinson’s disease with dementia, are likely provide insight.126 Some investigators are using PET to determine amyloid burden, integrity of the cholinergic system, and density of dopamine transporters in the neostriatum.2 Assays for CSF α-synuclein or plasma oligomers of α-synuclein in Parkinson’s disease remain in the discovery phase.127,128

BIOMARKERS FOR FRONTOTEMPORAL DEMENTIA

FTD comprises a constellation of neurodegenerative diseases that are currently under active investigation. There are relatively few studies of biomarkers for FTD, and all have focused on the dementia stage (little is known about preclinical or prodromal stages of this diagnostic group). One study observed that CSF T-tau and Aβ42 were changed similarly in FTD (n=34) and AD (n=76) when compared with controls (n=93), but that tau-P181 was significantly increased only in AD.129 Others could not confirm these findings for FTD, but did observe significantly increased neurofilament protein subtypes in a group of patients with FTD (n = 17) as compared with AD dementia (n = 20) or controls (n = 25).130 In relatively small sets of CSF samples, some groups have discovered potential protein profiles or MAPs that may prove useful in distinguishing patients with FTD from those with AD or controls.49,50,131 PET imaging showed that 11C-PiB retention in multiple cerebral regions, including frontal lobes, from patients with FTD was significantly lower than in patients with AD dementia and not significantly different from cognitively normal controls.54,132 As with CSF biomarkers, 20 to 33% of patients with FTD had 11C-PiB-PET signals that overlapped with values typical for patients with AD; however, this degree of overlap might diminish with improvements in clinical acumen.

CONCLUSIONS AND FUTURE DIRECTIONS

Dementing illnesses in elderly people pose an enormous challenge in the 21st century, and improved tools are urgently needed for diagnosis and monitoring of clinical progress and therapeutic interventions. Biomarker development holds considerable promise for these purposes, but the evidence base is much stronger at present for AD than for VCI and, in particular, for LBD and FTD. In general, the evidence is strongest for the dementia stages of these neurodegenerative diseases, weaker for prodromal illness, and weakest for preclinical stages. The immediate challenge will be to identify combinations of biomarkers that identify the relative contributions of AD, VCI, LBD, FTD, or other disease processes to the different stages of cognitive impairment in older individuals from the general population.31 The challenge will almost certainly be greatest for pre-clinical illness, for which therapies are most likely to be effective, but the tolerance for toxic effects is lowest.

Critical evaluation of the usefulness of a combination of standardized biomarkers, risk factor assessment, and structural and functional neuroimaging will be central areas of research as the field moves from discovery to validation to widespread clinical application. For example, will CSF tau (or a tau-P) and Aβ42 concentrations carry the same, substantially more, or less information than a PET imaging study of amyloid or glucose metabolism? The present challenge of contrasting performance of biomarkers in different studies could be resolved through a cooperative multi-center approach to the identification of markers most useful in widespread clinical application. We speculate that assessment of genetic risk, neuroimaging to determine structural or functional changes, and some ensemble of biomarkers will become part of the standard examination for geriatric dementia over the coming decade. New knowledge from these investigations will propel testing and development of mechanism-specific interventions through the identification of appropriate subgroups for clinical trials and provision of objective measures of disease suppression. Success with new therapies would lead quickly to the widespread application of biomarkers in the clinical management of geriatric cognitive disorders.

Acknowledgements

This work was supported by the Nancy and Buster Alvord Endowment and grants from the NIH (AG05136, AG08017, AG24010, AG24011, RR024140) and the Department of Veterans Affairs.

Search strategy and selection criteria:

References for this review were identified by searches of PubMed from 1966 until April 2008. Terms used for search included: biomarkers; Alzheimer's disease, mild cognitive impairment; vascular brain injury; vascular cognitive impairment; vascular dementia; Lewy body; and fronto-temporal dementia. Articles were also identified through searches of the authors’ own files. Only papers published in English were reviewed.

Contributions:

All authors were involved in the selection of papers to be included in this review and in writing and editing of subsequent drafts.

Footnotes

Conflicts of Interest:

There are no conflicts of interest for any author.

References

  • 1.Selnes OA, Vinters HV. Vascular cognitive impairment. Nat Clin Pract Neurol. 2006;2:538–547. doi: 10.1038/ncpneuro0294. [DOI] [PubMed] [Google Scholar]
  • 2.Lippa CF, Duda JE, Grossman M, Hurtig HI, Aarsland D, Boeve BF, et al. DLB and PDD boundary issues: diagnosis, treatment, molecular pathology, and biomarkers. Neurology. 2007;68:812–819. doi: 10.1212/01.wnl.0000256715.13907.d3. [DOI] [PubMed] [Google Scholar]
  • 3.Forman MS, Farmer J, Johnson JK, Clark CM, Arnold SE, Coslett HB, et al. Frontotemporal dementia: clinicopathological correlations. Ann Neurol. 2006;59:952–962. doi: 10.1002/ana.20873. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Sonnen JA, Larson EB, Crane PK, Haneuse S, Li G, Schellenberg GD, et al. Pathological correlates of dementia in a longitudinal, population-based sample of aging. Ann Neurol. 2007;62:406–413. doi: 10.1002/ana.21208. [DOI] [PubMed] [Google Scholar]
  • 5.White L, Petrovitch H, Hardman J, Nelson J, Davis D, Ross G, et al. Cerebrovascular pathology and dementia in autopsied Honolulu-Asia Aging Study participants. Ann N Y Acad Sci. 2002;977:9–23. doi: 10.1111/j.1749-6632.2002.tb04794.x. [DOI] [PubMed] [Google Scholar]
  • 6.Xuereb JH, Brayne C, Dufouil C, Gertz H, Wischik C, Harrington C, et al. Neuropathological findings in the very old. Results from the first 101 brains of a population-based longitudinal study of dementing disorders. Ann N Y Acad Sci. 2000;903:490–496. doi: 10.1111/j.1749-6632.2000.tb06404.x. [DOI] [PubMed] [Google Scholar]
  • 7.Schneider JA, Wilson RS, Bienias JL, Evans DA, Bennett DA. Cerebral infarctions and the likelihood of dementia from Alzheimer disease pathology. Neurology. 2004;62:1148–1155. doi: 10.1212/01.wnl.0000118211.78503.f5. [DOI] [PubMed] [Google Scholar]
  • 8.Bennett DA, Schneider JA, Bienias JL, Evans DA, Wilson RS. Mild cognitive impairment is related to Alzheimer disease pathology and cerebral infarctions. Neurology. 2005;64:834–841. doi: 10.1212/01.WNL.0000152982.47274.9E. [DOI] [PubMed] [Google Scholar]
  • 9.Neuropathology Group of the Medical Research Council Cognitive Function and Ageing Study (MRC CFAS) Pathological correlates of late-onset dementia in a multicentre, community-based population in England and Wales. Lancet. 2001;357:169–175. doi: 10.1016/s0140-6736(00)03589-3. [DOI] [PubMed] [Google Scholar]
  • 10.Rochet JC. Novel therapeutic strategies for the treatment of protein-misfolding diseases. Expert Rev Mol Med. 2007;9:1–34. doi: 10.1017/S1462399407000385. [DOI] [PubMed] [Google Scholar]
  • 11.Kwong LK, Neumann M, Sampathu DM, Lee VM, Trojanowski JQ. TDP-43 proteinopathy: the neuropathology underlying major forms of sporadic and familial frontotemporal lobar degeneration and motor neuron disease. Acta Neuropathol. 2007;114:63–70. doi: 10.1007/s00401-007-0226-5. [DOI] [PubMed] [Google Scholar]
  • 12.Katzman R. Editorial: The prevalence and malignancy of Alzheimer disease. A major killer. Arch Neurol. 1976;33:217–218. doi: 10.1001/archneur.1976.00500040001001. [DOI] [PubMed] [Google Scholar]
  • 13.Gatz M, Reynolds CA, Fratiglioni L, Johansson B, Mortimer JA, Berg S, et al. Role of genes and environments for explaining Alzheimer disease. Arch Gen Psychiatry. 2006;63:168–174. doi: 10.1001/archpsyc.63.2.168. [DOI] [PubMed] [Google Scholar]
  • 14.Riekse RG, Li G, Petrie EC, Leverenz JB, Vavrek D, Vuletic S, et al. Effect of statins on Alzheimer's disease biomarkers in cerebrospinal fluid. J Alzheimers Dis. 2006;10:399–406. doi: 10.3233/jad-2006-10408. [DOI] [PubMed] [Google Scholar]
  • 15.Quinn JF, Montine KS, Moore M, Morrow JD, Kaye JA, Montine TJ. Suppression of longitudinal increase in CSFF2-isoprostanes in Alzheimer's disease. Journal of Alzheimers Disease. 2004;6:93–97. doi: 10.3233/jad-2004-6110. [DOI] [PubMed] [Google Scholar]
  • 16.Peskind ER, Li G, Shofer J, Quinn JF, Kaye JA, Clark CM, et al. Age and apolipoprotein E*4 allele effects on cerebrospinal fluid beta-amyloid 42 in adults with normal cognition. Arch Neurol. 2006;63:936–939. doi: 10.1001/archneur.63.7.936. [DOI] [PubMed] [Google Scholar]
  • 17.Kauwe JS, Jacquart S, Chakraverty S, Wang J, Mayo K, Fagan AM, et al. Extreme cerebrospinal fluid amyloid beta levels identify family with late-onset Alzheimer's disease presenilin 1 mutation. Ann Neurol. 2007;61:446–453. doi: 10.1002/ana.21099. [DOI] [PubMed] [Google Scholar]
  • 18.Davis DG, Schmitt FA, Wekstein DR, Markesbery W. Alzheimer neuropathological alterations in aged cognitively normal subjects. J Neuropathol Exp Neurol. 1999;58:376–388. doi: 10.1097/00005072-199904000-00008. [DOI] [PubMed] [Google Scholar]
  • 19.Roman GC, Sachdev P, Royall DR, Bullock RA, Orgogozo JM, Lopez-Pousa S, et al. Vascular cognitive disorder: a new diagnostic category updating vascular cognitive impairment and vascular dementia. J Neurol Sci. 2004;226:81–87. doi: 10.1016/j.jns.2004.09.016. [DOI] [PubMed] [Google Scholar]
  • 20.Winblad B, Palmer K, Kivipelto M, Jelic V, Fratiglioni L, Wahlund LO, et al. Mild cognitive impairment--beyond controversies, towards a consensus: report of the International Working Group on Mild Cognitive Impairment. J Intern Med. 2004;256:240–246. doi: 10.1111/j.1365-2796.2004.01380.x. [DOI] [PubMed] [Google Scholar]
  • 21.McKhann G, Drachman D, Folstein M, Katzman R, Price D, Stadlan EM. Clinical diagnosis of Alzheimer's disease: report of the NINCDS-ADRDA Work Group under the auspices of the Department of Health and Human Services Task Force on Alzheimer's Disease. Neurology. 1984;34 doi: 10.1212/wnl.34.7.939. 939 – 344. [DOI] [PubMed] [Google Scholar]
  • 22.Hachinski V, Iadecola C, Petersen RC, Breteler MM, Nyenhuis DL, Black SE, et al. National Institute of Neurological Disorders and Stroke-Canadian Stroke Network vascular cognitive impairment harmonization standards. Stroke. 2006;37:2220–2241. doi: 10.1161/01.STR.0000237236.88823.47. [DOI] [PubMed] [Google Scholar]
  • 23.McKeith IG, Dickson DW, Lowe J, Emre M, O'Brien JT, Feldman H, et al. Diagnosis and management of dementia with Lewy bodies: third report of the DLB Consortium. Neurology. 2005;65:1863–1872. doi: 10.1212/01.wnl.0000187889.17253.b1. [DOI] [PubMed] [Google Scholar]
  • 24.Morris JC. Clinical dementia rating: a reliable and valid diagnostic and staging measure for dementia of the Alzheimer type. Int Psychogeriatr. 1997;9 Suppl 1:173–176. doi: 10.1017/s1041610297004870. discussion 7–8. [DOI] [PubMed] [Google Scholar]
  • 25.Petersen RC. The current status of mild cognitive impairment--what do we tell our patients? Nat Clin Pract Neurol. 2007;3:60–61. doi: 10.1038/ncpneuro0402. [DOI] [PubMed] [Google Scholar]
  • 26.Khachaturian AS, Corcoran CD, Mayer LS, Zandi PP, Breitner JC. Apolipoprotein E epsilon4 count affects age at onset of Alzheimer disease, but not lifetime susceptibility: The Cache County Study. Arch Gen Psychiatry. 2004;61:518–524. doi: 10.1001/archpsyc.61.5.518. [DOI] [PubMed] [Google Scholar]
  • 27.Sunderland T, Mirza N, Putnam KT, Linker G, Bhupali D, Durham R, et al. Cerebrospinal fluid beta-amyloid1-42 and tau in control subjects at risk for Alzheimer's disease: the effect of APOE epsilon4 allele. Biol Psychiatry. 2004;56:670–676. doi: 10.1016/j.biopsych.2004.07.021. [DOI] [PubMed] [Google Scholar]
  • 28.Fagan AM, Roe CM, Xiong C, Mintun MA, Morris JC, Holtzman DM. Cerebrospinal Fluid tau/beta-Amyloid42 Ratio as a Prediction of Cognitive Decline in Nondemented Older Adults. Arch Neurol. 2007;64:343–349. doi: 10.1001/archneur.64.3.noc60123. [DOI] [PubMed] [Google Scholar]
  • 29.Lewczuk P, Kornhuber J, Vanderstichele H, Vanmechelen E, Esselmann H, Bibl M, et al. Multiplexed quantification of dementia biomarkers in the CSF of patients with early dementias and MCI: a multicenter study. Neurobiol Aging. 2008;29:812–818. doi: 10.1016/j.neurobiolaging.2006.12.010. [DOI] [PubMed] [Google Scholar]
  • 30.Sunderland T, Linker G, Mirza N, Putnam K, Friedman D, Kimmel L, et al. Decreased beta-amyloid1-42 and increased tau levels in cerebrospinal fluid of patients with Alzheimer disease. Jama. 2003;289:2094–2103. doi: 10.1001/jama.289.16.2094. [DOI] [PubMed] [Google Scholar]
  • 31.Andreasen N, Minthon L, Davidsson P, Vanmechelen E, Vanderstichele H, Winblad B, et al. Evaluation of CSF-tau and CSF-Abeta42 as diagnostic markers for Alzheimer disease in clinical practice. Arch Neurol. 2001;58:373–379. doi: 10.1001/archneur.58.3.373. [DOI] [PubMed] [Google Scholar]
  • 32.Andreasen N, Vanmechelen E, Van de Voorde A, Davidsson P, Hesse C, Tarvonen S, et al. Cerebrospinal fluid tau protein as a biochemical marker for Alzheimer's disease: a community based follow up study. J Neurol Neurosurg Psychiatry. 1998;64:298–305. doi: 10.1136/jnnp.64.3.298. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Galasko D, Chang L, Motter R, Clark CM, Kaye J, Knopman D, et al. High cerebrospinal fluid tau and low amyloid beta42 levels in the clinical diagnosis of Alzheimer's disease and relation to apolipoprotein E genotype. Arch Neurol. 1998;55:937–945. doi: 10.1001/archneur.55.7.937. [DOI] [PubMed] [Google Scholar]
  • 34.Bouwman FH, Schoonenboom NS, Verwey NA, van Elk EJ, Kok A, Blankenstein MA, et al. CSF biomarker levels in early and late onset Alzheimer's disease. Neurobiol Aging. 2008 doi: 10.1016/j.neurobiolaging.2008.02.007. [DOI] [PubMed] [Google Scholar]
  • 35.Hansson O, Zetterberg H, Buchhave P, Andreasson U, Londos E, Minthon L, et al. Prediction of Alzheimer's disease using the CSF A beta 42/A beta 40 ratio in patients with mild cognitive impairment. Dementia and Geriatric Cognitive Disorders. 2007;23:316–320. doi: 10.1159/000100926. [DOI] [PubMed] [Google Scholar]
  • 36.Ewers M, Buerger K, Teipel SJ, Scheltens P, Schroder J, Zinkowski RP, et al. Multicenter assessment of CSF-phosphorylated tau for the prediction of conversion of MCI. Neurology. 2007;69:2205–2212. doi: 10.1212/01.wnl.0000286944.22262.ff. [DOI] [PubMed] [Google Scholar]
  • 37.Glodzik-Sobanska L, Pirraglia E, Brys M, de Santi S, Mosconi L, Rich KE, et al. The effects of normal aging and ApoE genotype on the levels of CSF biomarkers for Alzheimer's disease. Neurobiol Aging. 2007 doi: 10.1016/j.neurobiolaging.2007.08.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Buerger K, Teipel SJ, Zinkowski R, Sunderland T, Andreasen N, Blennow K, et al. Increased levels of CSF phosphorylated tau in apolipoprotein E epsilon4 carriers with mild cognitive impairment. Neurosci Lett. 2005;391:48–50. doi: 10.1016/j.neulet.2005.08.030. [DOI] [PubMed] [Google Scholar]
  • 39.Buerger K, Zinkowski R, Teipel SJ, Tapiola T, Arai H, Blennow K, et al. Differential diagnosis of Alzheimer disease with cerebrospinal fluid levels of tau protein phosphorylated at threonine 231. Arch Neurol. 2002;59:1267–1272. doi: 10.1001/archneur.59.8.1267. [DOI] [PubMed] [Google Scholar]
  • 40.Pratico D, Clark CM, Liun F, Rokach J, Lee VY, Trojanowski JQ. Increase of brain oxidative stress in mild cognitive impairment: a possible predictor of Alzheimer disease. Arch Neurol. 2002;59:972–976. doi: 10.1001/archneur.59.6.972. [DOI] [PubMed] [Google Scholar]
  • 41.de Leon MJ, Mosconi L, Blennow K, DeSanti S, Zinkowski R, Mehta PD, et al. Imaging and CSF studies in the preclinical diagnosis of Alzheimer's disease. Ann N Y Acad Sci. 2007;1097:114–145. doi: 10.1196/annals.1379.012. [DOI] [PubMed] [Google Scholar]
  • 42.Montine TJ, Beal MF, Cudkowicz ME, Brown RH, O'Donnell H, Margolin RA, et al. Increased cerebrospinal fluid F2-isoprostane concentration in probable Alzheimer's disease. Neurology. 1999;52:562–565. doi: 10.1212/wnl.52.3.562. [DOI] [PubMed] [Google Scholar]
  • 43.Montine TJ, Beal MF, Robertson D, Cudkowicz ME, Biaggioni I, O'Donnell H, et al. Cerebrospinal fluid F2-isoprostanes are elevated in Huntington's disease. Neurology. 1999;52:1104–1105. doi: 10.1212/wnl.52.5.1104. [DOI] [PubMed] [Google Scholar]
  • 44.Montine TJ, Sidell KR, Crews BC, Markesbery WR, Marnett LJ, Roberts LJ, 2nd, et al. Elevated CSF prostaglandin E2 levels in patients with probable AD. Neurology. 1999;53:1495–1498. doi: 10.1212/wnl.53.7.1495. [DOI] [PubMed] [Google Scholar]
  • 45.Montine TJ, Kaye JA, Montine KS, McFarland L, Morrow JD, Quinn JF. Cerebrospinal fluid A beta(42), tau, and F-2-isoprostane concentrations in patients with Alzheimer disease, other dementias, and in age-matched controls. Archives of Pathology & Laboratory Medicine. 2001;125:510–512. doi: 10.5858/2001-125-0510-CFATAF. [DOI] [PubMed] [Google Scholar]
  • 46.Pratico D, Clark CM, Lee VM, Trojanowski JQ, Rokach J, FitzGerald GA. Increased 8,12-iso-iPF2alpha-VI in Alzheimer's disease: correlation of a noninvasive index of lipid peroxidation with disease severity. Ann Neurol. 2000;48:809–812. [PubMed] [Google Scholar]
  • 47.de Leon MJ, DeSanti S, Zinkowski R, Mehta PD, Pratico D, Segal S, et al. Longitudinal CSF and MRI biomarkers improve the diagnosis of mild cognitive impairment. Neurobiol Aging. 2006;27:394–401. doi: 10.1016/j.neurobiolaging.2005.07.003. [DOI] [PubMed] [Google Scholar]
  • 48.Hu Y, Hosseini A, Kauwe JS, Gross J, Cairns NJ, Goate AM, et al. Identification and validation of novel CSF biomarkers for early stages of Alzheimer's disease. Proteomics Clin Appl. 2007;1:1373–1384. doi: 10.1002/prca.200600999. [DOI] [PubMed] [Google Scholar]
  • 49.Zhang J, Sokal I, Peskind ER, Quinn JF, Jankovic J, Kenney C, et al. CSF multianalyte profile distinguishes Alzheimer and Parkinson diseases. Am J Clin Pathol. 2008;129:526–529. doi: 10.1309/W01Y0B808EMEH12L. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Simonsen AH, McGuire J, Podust VN, Hagnelius NO, Nilsson TK, Kapaki E, et al. A novel panel of cerebrospinal fluid biomarkers for the differential diagnosis of Alzheimer's disease versus normal aging and frontotemporal dementia. Dement Geriatr Cogn Disord. 2007;24:434–440. doi: 10.1159/000110576. [DOI] [PubMed] [Google Scholar]
  • 51.Mintun MA, Larossa GN, Sheline YI, Dence CS, Lee SY, Mach RH, et al. [11C]PIB in a nondemented population: potential antecedent marker of Alzheimer disease. Neurology. 2006;67:446–452. doi: 10.1212/01.wnl.0000228230.26044.a4. [DOI] [PubMed] [Google Scholar]
  • 52.Pike KE, Savage G, Villemagne VL, Ng S, Moss SA, Maruff P, et al. Beta-amyloid imaging and memory in non-demented individuals: evidence for preclinical Alzheimer's disease. Brain. 2007;130:2837–2844. doi: 10.1093/brain/awm238. [DOI] [PubMed] [Google Scholar]
  • 53.Fagan AM, Mintun MA, Mach RH, Lee SY, Dence CS, Shah AR, et al. Inverse relation between in vivo amyloid imaging load and cerebrospinal fluid Abeta42 in humans. Ann Neurol. 2006;59:512–519. doi: 10.1002/ana.20730. [DOI] [PubMed] [Google Scholar]
  • 54.Rabinovici GD, Furst AJ, O'Neil JP, Racine CA, Mormino EC, Baker SL, et al. 11C-PIB PET imaging in Alzheimer disease and frontotemporal lobar degeneration. Neurology. 2007;68:1205–1212. doi: 10.1212/01.wnl.0000259035.98480.ed. [DOI] [PubMed] [Google Scholar]
  • 55.Archer HA, Edison P, Brooks DJ, Barnes J, Frost C, Yeatman T, et al. Amyloid load and cerebral atrophy in Alzheimer's disease: an 11C-PIB positron emission tomography study. Ann Neurol. 2006;60:145–147. doi: 10.1002/ana.20889. [DOI] [PubMed] [Google Scholar]
  • 56.Edison P, Archer HA, Hinz R, Hammers A, Pavese N, Tai YF, et al. Amyloid, hypometabolism, and cognition in Alzheimer disease: an [11C]PIB and [18F]FDG PET study. Neurology. 2007;68:501–508. doi: 10.1212/01.wnl.0000244749.20056.d4. [DOI] [PubMed] [Google Scholar]
  • 57.Klunk WE, Engler H, Nordberg A, Wang Y, Blomqvist G, Holt DP, et al. Imaging brain amyloid in Alzheimer's disease with Pittsburgh Compound-B. Ann Neurol. 2004;55:306–319. doi: 10.1002/ana.20009. [DOI] [PubMed] [Google Scholar]
  • 58.van Oijen M, Hofman A, Soares HD, Koudstaal PJ, Breteler MM. Plasma Abeta(1–40) and Abeta(1–42) and the risk of dementia: a prospective case-cohort study. Lancet Neurol. 2006;5:655–660. doi: 10.1016/S1474-4422(06)70501-4. [DOI] [PubMed] [Google Scholar]
  • 59.Graff-Radford NR, Crook JE, Lucas J, Boeve BF, Knopman DS, Ivnik RJ, et al. Association of low plasma abeta42/abeta40 ratios with increased imminent risk for mild cognitive impairment and Alzheimer disease. Arch Neurol. 2007;64:354–362. doi: 10.1001/archneur.64.3.354. [DOI] [PubMed] [Google Scholar]
  • 60.Sundelof J, Giedraitis V, Irizarry MC, Sundstrom J, Ingelsson E, Ronnemaa E, et al. Plasma beta amyloid and the risk of Alzheimer disease and dementia in elderly men: a prospective, population-based cohort study. Arch Neurol. 2008;65:256–263. doi: 10.1001/archneurol.2007.57. [DOI] [PubMed] [Google Scholar]
  • 61.Lopez OL, Kuller LH, Mehta PD, Becker JT, Gach HM, Sweet RA, et al. Plasma amyloid levels and the risk of AD in normal subjects in the Cardiovascular Health Study. Neurology. 2008;70:1664–1671. doi: 10.1212/01.wnl.0000306696.82017.66. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Maccioni RB, Lavados M, Guillon M, Mujica C, Bosch R, Farias G, et al. Anomalously phosphorylated tau and Abeta fragments in the CSF correlates with cognitive impairment in MCI subjects. Neurobiol Aging. 2006;27:237–244. doi: 10.1016/j.neurobiolaging.2005.01.011. [DOI] [PubMed] [Google Scholar]
  • 63.Bateman RJ, Wen G, Morris JC, Holtzman DM. Fluctuations of CSF amyloid-beta levels: implications for a diagnostic and therapeutic biomarker. Neurology. 2007;68:666–669. doi: 10.1212/01.wnl.0000256043.50901.e3. [DOI] [PubMed] [Google Scholar]
  • 64.Clark C, Xie S, Chittams J, Ewbank D, Peskind E, Galasko D, et al. Cerebrospinal fluid tau and beta-amyloid: how well do these biomarkers reflect autopsy-confirmed dementia diagnoses? Arch Neurol. 2003;60:1696–1702. doi: 10.1001/archneur.60.12.1696. [DOI] [PubMed] [Google Scholar]
  • 65.Chui HC, Zarow C, Mack WJ, Ellis WG, Zheng L, Jagust WJ, et al. Cognitive impact of subcortical vascular and Alzheimer's disease pathology. Ann Neurol. 2006;60:677–687. doi: 10.1002/ana.21009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Abdi F, Quinn JF, Jankovic J, McIntosh M, Leverenz JB, Peskind E, et al. Detection of biomarkers with a multiplex quantitative proteomic platform in cerebrospinal fluid of patients with neurodegenerative disorders. J Alzheimers Dis. 2006;9:293–348. doi: 10.3233/jad-2006-9309. [DOI] [PubMed] [Google Scholar]
  • 67.Hulstaert F, Blennow K, Ivanoiu A, Schoonderwaldt HC, Riemenschneider M, De Deyn PP, et al. Improved discrimination of AD patients using beta-amyloid(1–42) and tau levels in CSF. Neurology. 1999;52:1555–1562. doi: 10.1212/wnl.52.8.1555. [DOI] [PubMed] [Google Scholar]
  • 68.Blennow K. Cerebrospinal fluid protein biomarkers for Alzheimer's disease. NeuroRx. 2004;1:213–225. doi: 10.1602/neurorx.1.2.213. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Andreasen N, Blennow K. CSF biomarkers for mild cognitive impairment and early Alzheimer's disease. Clin Neurol Neurosurg. 2005;107:165–173. doi: 10.1016/j.clineuro.2004.10.011. [DOI] [PubMed] [Google Scholar]
  • 70.Engelborghs S, Sleegers K, Cras P, Brouwers N, Serneels S, De Leenheir E, et al. No association of CSF biomarkers with APOEepsilon4, plaque and tangle burden in definite Alzheimer's disease. Brain. 2007;130:2320–2326. doi: 10.1093/brain/awm136. [DOI] [PubMed] [Google Scholar]
  • 71.Herukka SK, Helisalmi S, Hallikainen M, Tervo S, Soininen H, Pirttila T. CSF A beta 42, Tau and phosphorylated Tau, APOE epsilon 4 allele and MCI type in progressive MCI. Neurobiology of Aging. 2007;28:507–514. doi: 10.1016/j.neurobiolaging.2006.02.001. [DOI] [PubMed] [Google Scholar]
  • 72.Hansson O, Zetterberg H, Buchhave P, Londos E, Blennow K, Minthon L. Association between CSF biomarkers and incipient Alzheimer's disease in patients with mild cognitive impairment: a follow-up study. Lancet Neurol. 2006;5:228–234. doi: 10.1016/S1474-4422(06)70355-6. [DOI] [PubMed] [Google Scholar]
  • 73.Andreasen N, Vanmechelen E, Vanderstichele H, Davidsson P, Blennow K. Cerebrospinal fluid levels of total-tau, phospho-tau and A beta 42 predicts development of Alzheimer's disease in patients with mild cognitive impairment. Acta Neurol Scand Suppl. 2003;179:47–51. doi: 10.1034/j.1600-0404.107.s179.9.x. [DOI] [PubMed] [Google Scholar]
  • 74.Buerger K, Teipel SJ, Zinkowski R, Blennow K, Arai H, Engel R, et al. CSF tau protein phosphorylated at threonine 231 correlates with cognitive decline in MCI subjects. Neurology. 2002;59:627–629. doi: 10.1212/wnl.59.4.627. [DOI] [PubMed] [Google Scholar]
  • 75.Arai H, Ishiguro K, Ohno H, Moriyama M, Itoh N, Okamura N, et al. CSF phosphorylated tau protein and mild cognitive impairment: a prospective study. Exp Neurol. 2000;166:201–203. doi: 10.1006/exnr.2000.7501. [DOI] [PubMed] [Google Scholar]
  • 76.Andreasen N, Minthon L, Vanmechelen E, Vanderstichele H, Davidsson P, Winblad B, et al. Cerebrospinal fluid tau and Abeta42 as predictors of development of Alzheimer's disease in patients with mild cognitive impairment. Neurosci Lett. 1999;273:5–8. doi: 10.1016/s0304-3940(99)00617-5. [DOI] [PubMed] [Google Scholar]
  • 77.Hoglund K, Hansson O, Buchhave P, Zetterberg H, Lewczuk P, Londos E, et al. Prediction of Alzheimer's Disease Using a Cerebrospinal Fluid Pattern of C-Terminally Truncated beta-Amyloid Peptides. Neurodegener Dis. 2008 doi: 10.1159/000119457. [DOI] [PubMed] [Google Scholar]
  • 78.Bouwman FH, van der Flier WM, Schoonenboom NS, van Elk EJ, Kok A, Rijmen F, et al. Longitudinal changes of CSF biomarkers in memory clinic patients. Neurology. 2007;69:1006–1011. doi: 10.1212/01.wnl.0000271375.37131.04. [DOI] [PubMed] [Google Scholar]
  • 79.Li G, Sokal I, Quinn JF, Leverenz JB, Brodey M, Schellenberg GD, et al. CSF tau/Abeta42 ratio for increased risk of mild cognitive impairment: a follow-up study. Neurology. 2007;69:631–639. doi: 10.1212/01.wnl.0000267428.62582.aa. [DOI] [PubMed] [Google Scholar]
  • 80.Pratico D, Lee VM, Trojanowski JQ, Rokach J, Fitzgerald GA. Increased F2-isoprostanes in Alzheimer's disease: evidence for enhanced lipid peroxidation in vivo. Faseb J. 1998;12:1777–1784. doi: 10.1096/fasebj.12.15.1777. [DOI] [PubMed] [Google Scholar]
  • 81.Reich EE, Markesbery WR, Roberts LJ, Swift LL, Morrow JD, Montine TJ. Brain regional quantification of F-ring and D-/E-ring isoprostanes and neuroprostanes in Alzheimer's disease. American Journal of Pathology. 2001;158:293–297. doi: 10.1016/S0002-9440(10)63968-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Markesbery WR, Kryscio RJ, Lovell MA, Morrow JD. Lipid peroxidation is an early event in the brain in amnestic mild cognitive impairment. Ann Neurol. 2005;58:730–735. doi: 10.1002/ana.20629. [DOI] [PubMed] [Google Scholar]
  • 83.Finehout EJ, Franck Z, Choe LH, Relkin N, Lee KH. Cerebrospinal fluid proteomic biomarkers for Alzheimer's disease. Ann Neurol. 2007;61:120–129. doi: 10.1002/ana.21038. [DOI] [PubMed] [Google Scholar]
  • 84.Castano EM, Roher AE, Esh CL, Kokjohn TA, Beach T. Comparative proteomics of cerebrospinal fluid in neuropathologically-confirmed Alzheimer's disease and non-demented elderly subjects. Neurol Res. 2006;28:155–163. doi: 10.1179/016164106X98035. [DOI] [PubMed] [Google Scholar]
  • 85.Hu Y, Malone JP, Fagan AM, Townsend RR, Holtzman DM. Comparative proteomic analysis of intra- and interindividual variation in human cerebrospinal fluid. Mol Cell Proteomics. 2005;4:2000–2009. doi: 10.1074/mcp.M500207-MCP200. [DOI] [PubMed] [Google Scholar]
  • 86.Jung SM, Lee K, Lee JW, Namkoong H, Kim HK, Kim S, et al. Both plasma retinol-binding protein and haptoglobin precursor allele 1 in CSF: Candidate biomarkers for the progression of normal to mild cognitive impairment to Alzheimer's disease. Neurosci Lett. 2008;436:153–157. doi: 10.1016/j.neulet.2008.03.010. [DOI] [PubMed] [Google Scholar]
  • 87.Simonsen AH, McGuire J, Hansson O, Zetterberg H, Podust VN, Davies HA, et al. Novel panel of cerebrospinal fluid biomarkers for the prediction of progression to Alzheimer dementia in patients with mild cognitive impairment. Arch Neurol. 2007;64:366–370. doi: 10.1001/archneur.64.3.366. [DOI] [PubMed] [Google Scholar]
  • 88.Peskind ER, Riekse R, Quinn JF, Kaye J, Clark CM, Farlow MR, et al. Safety and acceptability of the research lumbar puncture. Alzheimer Dis Assoc Disord. 2005;19:220–225. doi: 10.1097/01.wad.0000194014.43575.fd. [DOI] [PubMed] [Google Scholar]
  • 89.Montine TJ, Montine KS, McMahan W, Markesbery WR, Quinn JF, Morrow JD. F-2-isoprostanes in Alzheimer and other neurodegenerative diseases. Antioxidants & Redox Signaling. 2005;7:269–275. doi: 10.1089/ars.2005.7.269. [DOI] [PubMed] [Google Scholar]
  • 90.Irizarry MC, Yao Y, Hyman BT, Growdon JH, Pratico D. Plasma F2A isoprostane levels in Alzheimer's and Parkinson's disease. Neurodegener Dis. 2007;4:403–405. doi: 10.1159/000107699. [DOI] [PubMed] [Google Scholar]
  • 91.Hardy J. Amyloid, the presenilins and Alzheimer's disease. Trends Neurosci. 1997;20:154–159. doi: 10.1016/s0166-2236(96)01030-2. [DOI] [PubMed] [Google Scholar]
  • 92.Hutton M, Hardy J. The presenilins and Alzheimer's disease. Hum Mol Genet. 1997;6:1639–1646. doi: 10.1093/hmg/6.10.1639. [DOI] [PubMed] [Google Scholar]
  • 93.Mehta PD, Capone G, Jewell A, Freedland RL. Increased amyloid beta protein levels in children and adolescents with Down syndrome. J Neurol Sci. 2007;254:22–27. doi: 10.1016/j.jns.2006.12.010. [DOI] [PubMed] [Google Scholar]
  • 94.Mayeux R, Honig LS, Tang MX, Manly J, Stern Y, Schupf N, et al. Plasma A[beta]40 and A[beta]42 and Alzheimer's disease: relation to age, mortality, and risk. Neurology. 2003;61:1185–1190. doi: 10.1212/01.wnl.0000091890.32140.8f. [DOI] [PubMed] [Google Scholar]
  • 95.Pomara N, Willoughby LM, Sidtis JJ, Mehta PD. Selective reductions in plasma Abeta 1–42 in healthy elderly subjects during longitudinal follow-up: a preliminary report. Am J Geriatr Psychiatry. 2005;13:914–917. doi: 10.1176/appi.ajgp.13.10.914. [DOI] [PubMed] [Google Scholar]
  • 96.Ertekin-Taner N, Graff-Radford N, Younkin LH, Eckman C, Adamson J, Schaid DJ, et al. Heritability of plasma amyloid beta in typical late-onset Alzheimer's disease pedigrees. Genet Epidemiol. 2001;21:19–30. doi: 10.1002/gepi.1015. [DOI] [PubMed] [Google Scholar]
  • 97.Blennow K, de Leon MJ, Zetterberg H. Alzheimer's disease. Lancet. 2006;368:387–403. doi: 10.1016/S0140-6736(06)69113-7. [DOI] [PubMed] [Google Scholar]
  • 98.Ciabattoni G, Porreca E, Di Febbo C, Di Iorio A, Paganelli R, Bucciarelli T, et al. Determinants of platelet activation in Alzheimer's disease. Neurobiol Aging. 2007;28:336–342. doi: 10.1016/j.neurobiolaging.2005.12.011. [DOI] [PubMed] [Google Scholar]
  • 99.Tan ZS, Beiser AS, Vasan RS, Roubenoff R, Dinarello CA, Harris TB, et al. Inflammatory markers and the risk of Alzheimer disease: the Framingham Study. Neurology. 2007;68:1902–1908. doi: 10.1212/01.wnl.0000263217.36439.da. [DOI] [PubMed] [Google Scholar]
  • 100.van Oijen M, Witteman JC, Hofman A, Koudstaal PJ, Breteler MM. Fibrinogen is associated with an increased risk of Alzheimer disease and vascular dementia. Stroke. 2005;36:2637–2641. doi: 10.1161/01.STR.0000189721.31432.26. [DOI] [PubMed] [Google Scholar]
  • 101.Sun YX, Minthon L, Wallmark A, Warkentin S, Blennow K, Janciauskiene S. Inflammatory markers in matched plasma and cerebrospinal fluid from patients with Alzheimer's disease. Dement Geriatr Cogn Disord. 2003;16:136–144. doi: 10.1159/000071001. [DOI] [PubMed] [Google Scholar]
  • 102.Rentzos M, Michalopoulou M, Nikolaou C, Cambouri C, Rombos A, Dimitrakopoulos A, et al. Serum levels of soluble intercellular adhesion molecule-1 and soluble endothelial leukocyte adhesion molecule-1 in Alzheimer's disease. J Geriatr Psychiatry Neurol. 2004;17:225–231. doi: 10.1177/0891988704269822. [DOI] [PubMed] [Google Scholar]
  • 103.Matsubara E, Hirai S, Amari M, Shoji M, Yamaguchi H, Okamoto K, et al. Alpha 1-antichymotrypsin as a possible biochemical marker for Alzheimer-type dementia. Ann Neurol. 1990;28:561–567. doi: 10.1002/ana.410280414. [DOI] [PubMed] [Google Scholar]
  • 104.Licastro F, Pedrini S, Caputo L, Annoni G, Davis LJ, Ferri C, et al. Increased plasma levels of interleukin-1, interleukin-6 and alpha-1-antichymotrypsin in patients with Alzheimer's disease: peripheral inflammation or signals from the brain? J Neuroimmunol. 2000;103:97–102. doi: 10.1016/s0165-5728(99)00226-x. [DOI] [PubMed] [Google Scholar]
  • 105.Pirttila T, Mehta PD, Frey H, Wisniewski HM. Alpha 1-antichymotrypsin and IL-1 beta are not increased in CSF or serum in Alzheimer's disease. Neurobiol Aging. 1994;15:313–317. doi: 10.1016/0197-4580(94)90026-4. [DOI] [PubMed] [Google Scholar]
  • 106.Ray S, Britschgi M, Herbert C, Takeda-Uchimura Y, Boxer A, Blennow K, et al. Classification and prediction of clinical Alzheimer's diagnosis based on plasma signaling proteins. Nat Med. 2007;13:1359–1362. doi: 10.1038/nm1653. [DOI] [PubMed] [Google Scholar]
  • 107.Johnson KA, Gregas M, Becker JA, Kinnecom C, Salat DH, Moran EK, et al. Imaging of amyloid burden and distribution in cerebral amyloid angiopathy. Ann Neurol. 2007;62:229–234. doi: 10.1002/ana.21164. [DOI] [PubMed] [Google Scholar]
  • 108.Bacskai BJ, Frosch MP, Freeman SH, Raymond SB, Augustinack JC, Johnson KA, et al. Molecular imaging with Pittsburgh Compound B confirmed at autopsy: a case report. Arch Neurol. 2007;64:431–434. doi: 10.1001/archneur.64.3.431. [DOI] [PubMed] [Google Scholar]
  • 109.Fodero-Tavoletti MT, Smith DP, McLean CA, Adlard PA, Barnham KJ, Foster LE, et al. In vitro characterization of Pittsburgh compound-B binding to Lewy bodies. J Neurosci. 2007;27:10365–10371. doi: 10.1523/JNEUROSCI.0630-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Boxer AL, Rabinovici GD, Kepe V, Goldman J, Furst AJ, Huang SC, et al. Amyloid imaging in distinguishing atypical prion disease from Alzheimer disease. Neurology. 2007;69:283–290. doi: 10.1212/01.wnl.0000265815.38958.b6. [DOI] [PubMed] [Google Scholar]
  • 111.Rowe CC, Ackerman U, Browne W, Mulligan R, Pike KL, O'Keefe G, et al. Imaging of amyloid beta in Alzheimer's disease with 18F-BAY94-9172, a novel PET tracer: proof of mechanism. Lancet Neurol. 2008;7:129–135. doi: 10.1016/S1474-4422(08)70001-2. [DOI] [PubMed] [Google Scholar]
  • 112.Reed BR, Mungas DM, Kramer JH, Ellis W, Vinters HV, Zarow C, et al. Profiles of neuropsychological impairment in autopsy-defined Alzheimer's disease and cerebrovascular disease. Brain. 2007;130:731–739. doi: 10.1093/brain/awl385. [DOI] [PubMed] [Google Scholar]
  • 113.Stefani A, Bernardini S, Panella M, Pierantozzi M, Nuccetelli M, Koch G, et al. AD with subcortical white matter lesions and vascular dementia: CSF markers for differential diagnosis. J Neurol Sci. 2005;237:83–88. doi: 10.1016/j.jns.2005.05.016. [DOI] [PubMed] [Google Scholar]
  • 114.Hampel H, Buerger K, Zinkowski R, Teipel SJ, Goernitz A, Andreasen N, et al. Measurement of phosphorylated tau epitopes in the differential diagnosis of Alzheimer disease: a comparative cerebrospinal fluid study. Arch Gen Psychiatry. 2004;61:95–102. doi: 10.1001/archpsyc.61.1.95. [DOI] [PubMed] [Google Scholar]
  • 115.Skoog I, Wallin A, Fredman P, Hesse C, Aevarsson O, Karlsson I, et al. A population study on blood-brain barrier function in 85-year-olds: relation to Alzheimer's disease and vascular dementia. Neurology. 1998;50:966–971. doi: 10.1212/wnl.50.4.966. [DOI] [PubMed] [Google Scholar]
  • 116.Wallin A, Sjogren M. Cerebrospinal fluid cytoskeleton proteins in patients with subcortical white-matter dementia. Mech Ageing Dev. 2001;122:1937–1949. doi: 10.1016/s0047-6374(01)00306-2. [DOI] [PubMed] [Google Scholar]
  • 117.Sjogren M, Blomberg M, Jonsson M, Wahlund LO, Edman A, Lind K, et al. Neurofilament protein in cerebrospinal fluid: a marker of white matter changes. J Neurosci Res. 2001;66:510–516. doi: 10.1002/jnr.1242. [DOI] [PubMed] [Google Scholar]
  • 118.Maruyama M, Matsui T, Tanji H, Nemoto M, Tomita N, Ootsuki M, et al. Cerebrospinal fluid tau protein and periventricular white matter lesions in patients with mild cognitive impairment: implications for 2 major pathways. Arch Neurol. 2004;61:716–720. doi: 10.1001/archneur.61.5.716. [DOI] [PubMed] [Google Scholar]
  • 119.Tullberg M, Mansson JE, Fredman P, Lekman A, Blennow K, Ekman R, et al. CSF sulfatide distinguishes between normal pressure hydrocephalus and subcortical arteriosclerotic encephalopathy. J Neurol Neurosurg Psychiatry. 2000;69:74–81. doi: 10.1136/jnnp.69.1.74. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Fredman P, Wallin A, Blennow K, Davidsson P, Gottfries CG, Svennerholm L. Sulfatide as a biochemical marker in cerebrospinal fluid of patients with vascular dementia. Acta Neurol Scand. 1992;85:103–106. doi: 10.1111/j.1600-0404.1992.tb04006.x. [DOI] [PubMed] [Google Scholar]
  • 121.Adair JC, Charlie J, Dencoff JE, Kaye JA, Quinn JF, Camicioli RM, et al. Measurement of gelatinase B (MMP-9) in the cerebrospinal fluid of patients with vascular dementia and Alzheimer disease. Stroke. 2004;35:e159–e162. doi: 10.1161/01.STR.0000127420.10990.76. [DOI] [PubMed] [Google Scholar]
  • 122.Bibl M, Esselmann H, Mollenhauer B, Weniger G, Welge V, Liess M, et al. Blood-based neurochemical diagnosis of vascular dementia: a pilot study. J Neurochem. 2007;103:467–474. doi: 10.1111/j.1471-4159.2007.04763.x. [DOI] [PubMed] [Google Scholar]
  • 123.Ravaglia G, Forti P, Maioli F, Chiappelli M, Montesi F, Tumini E, et al. Blood inflammatory markers and risk of dementia: The Conselice Study of Brain Aging. Neurobiol Aging. 2007;28:1810–1820. doi: 10.1016/j.neurobiolaging.2006.08.012. [DOI] [PubMed] [Google Scholar]
  • 124.Vanderstichele H, De Vreese K, Blennow K, Andreasen N, Sindic C, Ivanoiu A, et al. Analytical performance and clinical utility of the INNOTEST PHOSPHO-TAU181P assay for discrimination between Alzheimer's disease and dementia with Lewy bodies. Clin Chem Lab Med. 2006;44:1472–1480. doi: 10.1515/CCLM.2006.258. [DOI] [PubMed] [Google Scholar]
  • 125.Bibl M, Mollenhauer B, Esselmann H, Lewczuk P, Klafki HW, Sparbier K, et al. CSF amyloid-beta-peptides in Alzheimer's disease, dementia with Lewy bodies and Parkinson's disease dementia. Brain. 2006;129:1177–1187. doi: 10.1093/brain/awl063. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Parnetti L, Tiraboschi P, Lanari A, Peducci M, Padiglioni C, D'Amore C, et al. Cerebrospinal Fluid Biomarkers in Parkinson's Disease with Dementia and Dementia with Lewy Bodies. Biol Psychiatry. 2008 doi: 10.1016/j.biopsych.2008.02.016. [DOI] [PubMed] [Google Scholar]
  • 127.El-Agnaf OM, Salem SA, Paleologou KE, Curran MD, Gibson MJ, Court JA, et al. Detection of oligomeric forms of alpha-synuclein protein in human plasma as a potential biomarker for Parkinson's disease. Faseb J. 2006;20:419–425. doi: 10.1096/fj.03-1449com. [DOI] [PubMed] [Google Scholar]
  • 128.Tokuda T, Salem SA, Allsop D, Mizuno T, Nakagawa M, Qureshi MM, et al. Decreased alpha-synuclein in cerebrospinal fluid of aged individuals and subjects with Parkinson's disease. Biochem Biophys Res Commun. 2006;349:162–166. doi: 10.1016/j.bbrc.2006.08.024. [DOI] [PubMed] [Google Scholar]
  • 129.Kapaki E, Paraskevas GP, Papageorgiou SG, Bonakis A, Kalfakis N, Zalonis I, et al. Diagnostic Value of CSF Biomarker Profile in Frontotemporal Lobar Degeneration. Alzheimer Dis Assoc Disord. 2008;22:47–53. doi: 10.1097/WAD.0b013e3181610fea. [DOI] [PubMed] [Google Scholar]
  • 130.Pijnenburg YA, Janssen JC, Schoonenboom NS, Petzold A, Mulder C, Stigbrand T, et al. CSF neurofilaments in frontotemporal dementia compared with early onset Alzheimer's disease and controls. Dement Geriatr Cogn Disord. 2007;23:225–230. doi: 10.1159/000099473. [DOI] [PubMed] [Google Scholar]
  • 131.Ruetschi U, Zetterberg H, Podust VN, Gottfries J, Li S, Hviid Simonsen A, et al. Identification of CSF biomarkers for frontotemporal dementia using SELDI-TOF. Exp Neurol. 2005;196:273–281. doi: 10.1016/j.expneurol.2005.08.002. [DOI] [PubMed] [Google Scholar]
  • 132.Engler H, Santillo AF, Wang SX, Lindau M, Savitcheva I, Nordberg A, et al. In vivo amyloid imaging with PET in frontotemporal dementia. Eur J Nucl Med Mol Imaging. 2008;35:100–106. doi: 10.1007/s00259-007-0523-1. [DOI] [PubMed] [Google Scholar]

RESOURCES