Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2018 Oct 10.
Published in final edited form as: Am J Med Genet B Neuropsychiatr Genet. 2017 Jan;174(1):5–26. doi: 10.1002/ajmg.b.32499

Genomic Variants, Genes, and Pathways of Alzheimer’s Disease: An Overview

Adam C Naj 1, Gerard D Schellenberg 2; for the Alzheimer’s Disease Genetics Consortium (ADGC)
PMCID: PMC6179157  NIHMSID: NIHMS971364  PMID: 27943641

Abstract

Alzheimer’s disease (AD) (MIM: 104300) is a highly heritable disease with great complexity in its genetic contributors, and represents the most common form of dementia. With the gradual aging of the world’s population, leading to increased prevalence of AD, and the substantial cost of care for those afflicted, identifying the genetic causes of disease represents a critical effort in identifying therapeutic targets. Here we provide a comprehensive review of genomic studies of AD, from the earliest linkage studies identifying monogenic contributors to early-onset forms of AD to the genome-wide and rare variant association studies of recent years that are being used to characterize the mosaic of genetic contributors to late-onset AD (LOAD), and which have identified approximately ~20 genes with common variants contributing to LOAD risk. In addition, we explore studies employing alternative approaches to identify genetic contributors to AD, including studies of AD-related phenotypes and multi-variant association studies such as pathway analyses. Finally, we introduce studies of next-generation sequencing, which have recently helped identify multiple low-frequency and rare variant contributors to AD, and discuss on-going efforts with next-generation sequencing studies to develop statistically well-powered and comprehensive genomic studies of AD. Through this review, we help uncover the many insights the genetics of AD have provided into the pathways and pathophysiology of AD.

Keywords: Late-onset Alzheimer’s Disease (LOAD), Alzheimer’s Disease (AD), genome-wide association studies (GWAS), next-generation sequencing, pathway analysis

Alzheimer’s Disease: Characteristics, Epidemiology, and Genetic Contribution

Alzheimer’s disease (AD) is characterized by an insidious onset and slowly progressing memory loss, and advancing to deficits in higher intellectual functions and cognitive abilities, typically across multiple domains (language, praxis, recognition, or executive functioning). In the 1900s, the disease was identified in a dementia patient by Alois Alzheimer, who described the characteristic histopathologic signs of the disease, senile (neuritic) plaques and neurofibrillary tangles (Alzheimer, 1907). At that time, AD was thought to be a rare cause of senility, a notion that was refuted in the 1960s with an autopsy series by Blessed, Tomlinson, and Roth performed on hundreds of brains affected with ‘normal senility,’ which revealed that the majority were affected by the plaque and tangle lesions characteristic of AD (1968). Subsequent work revealed that most dementias arose from specific pathological processes and were not normal features of aging, and efforts to characterize pathology and correlate clinical manifestations culminated in the development of common criteria for clinical diagnosis in 1984 (McKhann et al.). Clinical diagnosis of AD is accurate in 85-90% of cases (Dal Forno et al., 1996; Pericak-Vance et al., 1995; Welsh-Bohmer et al., 1997), but remains a diagnosis of exclusion, complicated by multiple factors: cognitive testing revealing considerable overlap in cognitive function between demented and non-demented elderly (Knopman, 2003); presence of multiple types of psychiatric co-morbidities including depression (25-75%) (Lee and Lyketsos, 2003; Olin et al., 2002), anxiety (20-70%) (Porter et al., 2003; Teri et al., 1999), and psychosis (41.1%) (Ropacki and Jeste, 2005); and co-morbid dementias including Vascular Dementia (VaD; 38-58%) (Fernando et al., 2004; Schneider et al., 2007). Definitive diagnosis still requires post-mortem neuropathologic examination of brain tissues.

AD is the leading cause of dementia in the elderly and sixth leading cause of death in the United States, affecting 5.3 million people in the United States including approximately 200,000 individuals under age 65 (Alzheimer’s Association, 2015; Hebert et al., 2013). AD prevalence also increases with age, from 0.3-0.5% at age 60, to 11-15% at age 80 (Bachman et al., 1992; Prencipe et al., 1996; Rocca et al., 1991), although there is consider able variability in these rate estimates, with some reports finding prevalence as high as 47% after age 84 (Evans et al., 1989). Half of the beds in long-term care facilities are already devoted to patients with dementia, and a majority of those patients have AD (Katzman, 1986). As the population continues to age, the prevalence is expected to increase almost threefold by 2050 (Hebert et al., 2003), making AD not only a common, but also a growing health crisis.

Many lines of evidence point to a genetic basis for the development of AD. Familial aggregation studies demonstrated clustering of AD within families (Heyman et al., 1983; Nee et al., 1983; Sjogren et al., 1952), both among those with early-onset forms (St George-Hyslop et al., 1989) and the much more common late-onset form of AD (Pericak-Vance et al., 1988), where a more complex genetic etiology was suggested. Recurrence risk among siblings of those with AD (λS) have consistently ranged between four and five times higher than the risk to the general population (Hirst et al., 1994; Pericak-Vance et al., 1988; Sadovnick et al., 1989). Twin studies found higher concordance of AD among monozygotic twins (22-83%) than dizygotic twins (0-50%) (Bergem, 1994; Bergem et al., 1997; Breitner et al., 1993; Breitner et al., 1995), lending further support to genetic contributions to AD over only shared environmental contributors. Among early-onset families, patterns of AD inheritance have been consistent with classical Mendelian autosomal dominant inheritance, while the late-onset families have tended to suggest a multifactorial inheritance involving both genetic and non-genetic factors (Farrer et al., 1991; Rao et al., 1994; van Duijn et al., 1993). These studies provided powerful evidence for a genetic contribution to AD risk, even before any specific genetic variants were described.

Genetics of Alzheimer Disease: Early-Onset AD

In the early to mid-1990s, genetic studies of AD focused on extended families with high burden of disease (two or more cases among first-degree relatives), and used linkage analysis of highly polymorphic genetic markers called short tandem repeats (STRs, or microsattelites) in order to identify genomic regions co-transmitting with disease in affected family members (Table 1). This strategy, followed by “fine mapping”- the positional cloning of candidate genes- was used to identify genes and genetic variants contributing to AD risk. The first three genes known to cause AD were identified among families with multiple early-onset cases (age-at-onset < 60 years): APP, encoding amyloid precursor protein (Goate et al., 1991), and PS1 and PS2, encoding presenilins I and II respectively (Levy-Lahad et al., 1995; Rogaev et al., 1995; Sherrington et al., 1995), each transmitting disease-causing variants in the predicted autosomal-dominant fashion.

Table 1.

Genomic regions of linkage to Alzheimer’s disease observed in two or more independent linkage datasets.

Chr Region # Studies Studies LOD Scores Relevant Genes
1 1p13.3-q23.3 2 (Butler et al., 2009; Liu et al., 2007) 5.2 GSTM4; GSTM1; GSTM3; CSF1; NGF; HMGCS2; PRKAB2; APH1A; CTSS; THEM5; FAM63A; CHRNB2; LMNA; PMVK; FDPS; APOA1BP; GBA; NTRK1; CRP; NCSTN
1q23.3-q32.3 4 (Blacker et al., 2003; Butler et al., 2009; Kunkle et al., 2016; Liu et al., 2007) 2.1–4.0 F11R; USF1; FCER1G; RGS4; APOA2; RXRG; POU2F1; PRDX6; SOAT1; PTGS2; CR1
3 3q12.3-q27.3 4 (Barral et al., 2015; Butler et al., 2009; Cummings et al., 2012; Kunkle et al., 2016) 3.5–4.5 DRD3; GSK3B; TF; MME; BCHE; SLC2A2; AHSG; SST
5 5p13-p15 8 (Barral et al., 2015; Blacker et al., 2003; Curtis et al., 2001; Kunkle et al., 2016; Lee et al., 2006; Myers et al., 2002; Olson et al., 2002; Pericak-Vance et al., 2000) 1.4–2.8 SLC6A3; PRKAA1
8 8p22-p21.1 2 (Butler et al., 2009; Lee et al., 2008) > 2.0 NAT1; NAT2; LPL; ADRA1A; CHRNA2; CLU
8q22.3 2 (Cummings et al., 2012; Kunkle et al., 2016) 0.9-3.6 DPYS
9 9p21-9p22.1 6 (Curtis et al., 2001; Hamshere et al., 2007; Kunkle et al., 2016; Myers et al., 2002; Pericak-Vance et al., 2000; Scott et al., 2003) > 1.0–4.6 IFT74
9q22-q34 9 (Barral et al., 2015; Blacker et al., 2003; Curtis et al., 2001; Hamshere et al., 2007; Holmans et al., 2005; Lee et al., 2006; Myers et al., 2002; Olson et al., 2002; Pericak-Vance et al., 2000) 1.6–4.2 FBP1; GOLM1; ABCA1; DFNB31; TLR4; NDUFA8; PSMB7; HSPA5; POMT1; DBH; RXRA; TRAF2; ABCA2
10 10q21-q22 7 (Blacker et al., 2003; Curtis et al., 2001; Hamshere et al., 2007; Holmans et al., 2005; Liu et al., 2007; Myers et al., 2002; Olson et al., 2002) 1.8–4.2 ZWINT; UBE2D1; TFAM; BICC1; ANK3; CDC2; EGR2; CTNNA3; LRRTM3; DNAJC12; SIRT1; SRGN; SUPV3L1; TSPAN15; VPS26A; HK1; TACR2; NEUROG3; SAR1A; SGPL1; PSAP; CHST3; PPP3CB; SEC24C; NDST2; CAMK2G; PLAU; VCL; AP3M1; MYST4; KCNMA1
11 11q12.1-q14.1 2 (Barral et al., 2015; Kunkle et al., 2016) 1.0–4.7 CNTF; CHRM1; RTN3; GSTP1; GAL; FADD; INPPL1; UCP2; GAB2; PICALM
12 12p11-p13 4 (Curtis et al., 2001; Holmans et al., 2005; Myers et al., 2002; Pericak-Vance et al., 1997) 1.4–3.9 TNFRSF1A; CNAP1; GAPDH; GNB3; C1R; APOBEC1; MMP3; A2M; PZP; A2MP; OLR1; LRP6; GRIN2B; GYS2; ABCC9; PKP2P1
14 14q12-q13.3 2 (Barral et al., 2015; Kunkle et al., 2016) 2.1-3.9 CTSG; NFKBIA
19 19q12-q13.33 10 (Blacker et al., 2003; Butler et al., 2009; Curtis et al., 2001; Hamshere et al., 2007; Holmans et al., 2005; Kunkle et al., 2016; Li et al., 2002; Myers et al., 2002; Olson et al., 2002; Pericak-Vance et al., 2000; Sillen et al., 2008) 1.6–7.7 LRP3; USF2; GAPDHS; PSENEN; AKT2; TGFB1; LIPE; XRCC1; BCL3; APOE; PVRL2; TOMM40; APOC1; APOC2; ERCC2; CARD8; GYS1; LHB; CD33; NR1H2
21 21q21-q22 4 (Blacker et al., 2003; Holmans et al., 2005; Myers et al., 2002; Olson et al., 2002) 1.6–4.5 PRSS7; NCAM2; APP; C21ORF63; C21ORF55; RUNX1; DYRK1A; KCNJ6; BACE2

Notes: Chr = chromosome; Bolded gene names are known GWAS signals with genome-wide statistical significance

APP, identified in 1987 (Goldgaber et al., 1987; Tanzi et al., 1987), was initially cloned and localized to chromosome 21. Down Syndrome (Trisomy-21; OMIM: 190685), which involves whole or partial duplication of chromosome 21, is the most common cause of mental retardation, and persons with Trisomy-21 often develop the neuritic plaques and neurofibrillary tangles characteristic of AD at an early age (Lemere et al., 1996). Further confirmation of the role of APP was revealed through mutational analysis of an individual with Down Syndrome who was free of Alzheimer’s type dementia and whose autopsy showed no neuritic plaque or neurofibrillary tangle development in the brain; the individual carried only a partial duplication of chromosome 21 which excluded duplication of APP (Prasher et al., 1998). Functional studies demonstrated the role of mutations in this gene appear to cause shifts in the proteolytic cleavage of APP toward amyloidogenic pathways, leading to accumulation of an isoform of beta-amyloid Aβ42, which is less soluble than the common isoform Aβ40 (Suzuki et al., 1994), and is neurotoxic (Hilbich et al., 1991).

Not all families with early-onset AD showed linkage signals on chromosome 21 or were affected by these mutations (Pericak-Vance et al., 1988; Schellenberg et al., 1987), and subsequent studies identified linkage to chromosome 14 (Schellenberg et al., 1992; St George-Hyslop et al., 1992; Van Broeckhoven et al., 1994). In 1995, AD-causing mutations were identified in PSEN1 (Sherrington et al., 1995), which contributed to an aggressive form of AD (Selkoe, 2001) with relative abundance of Aβ42 accumulation (Lemere et al., 1996; Mann et al., 1996). Even after PSEN1 had been identified, there remained families with heavy burden early-onset AD who demonstrated linkage to neither chromosomes 21 nor 14. Many of these families shared common ancestral origins in the Volga river basin of Russia (Bird et al., 1988), and showed linkage to chromosome 1 (Levy-Lahad et al., 1995). A large 7.5 kb alternative polyadenylation message was identified when cloning one gene in the linkage region, and the gene identified demonstrated a high degree of homology to PSEN1, and was thus identified as PSEN2. Mutations in PSEN2 are rare, but have been described in the Volga German families and in one Italian pedigree (Levy-Lahad et al., 1995; Rogaev et al., 1995). Individuals affected by PSEN2 mutations have highly variable ages of disease onset, as early as their 40s and as late as their 80s. A surfeit of evidence suggest that presenilins contribute to a partial loss of function in the γ-secretase complex, which affects several downstream signalling pathways, and increase the vulnerability of the brain to Aβ toxicity(De Strooper, 2007), however the precise roles of presenilin mutations continue to be investigated.

Genetics of Alzheimer Disease: Late-Onset Alzheimer Disease (LOAD)

LOAD cases comprise the majority (90-95% or more) of individuals affected with AD, and as noted earlier, there is considerable evidence to support a strong genetic component in the etiology of LOAD, including a greater lifetime risk for dementia among offspring in late-onset families compared to offspring in early-onset families (Farrer et al., 1990). Identifying genetic contributors to LOAD has posed great challenges, as while EOAD is characterized by highly penetrant mutations in a few known risk genes, LOAD is likely caused by multiple low penetrance genetic variants.

The first susceptibility locus for LOAD: APOE

The first confirmed susceptibility gene for LOAD, the Apolipoprotein E (APOE) gene, was initially discovered using linkage analysis (Pericak-Vance et al., 1991) in a subset of AD affected individuals with an age-at-onset (AAO) greater than or equal to 60 years of age, followed by association analysis (Strittmatter et al., 1993). Located on chromosome 19 (19q13.2) (Das et al., 1985; Olaisen et al., 1982), APOE encodes apolipoprotein E (ApoE), a serum protein synthesized by astrocytes (Mahley, 1988) involved in the transport, storage, and metabolism of lipid. The three known isoforms of the ApoE protein (ApoE2, ApoE3, and ApoE4) result from single amino acid substitutions at residues 112 and 158 (Rall et al., 1982; Weisgraber et al., 1982), and the corresponding encoding, ε2, ε3, and ε4, are estimated to be 8%, 78%, and 14%, respectively, in most Caucasian populations (Martins et al., 1995), while other ApoE isoforms (such as those encoded by ε1, ε5, and ε7) are extremely rare. Compared to the common ε3 allele, the ε4 allele has consistently been found to be positively associated with risk for sporadic and familial AD, whereas the ε2 allele appears to be a protective factor for risk of LOAD (Martins et al., 1995). The ε4 allele also has a dose-dependent relationship with LOAD susceptibility (Corder et al., 1994; Frisoni et al., 1995) with the relative risks of LOAD for ε4 homozygotes and for ε3/ε4 heterozygotes are approximately 15 and 3, respectively, compared to ε3 homozygotes (Farrer et al., 1997). The APOE ε2/ε3/ε4 polymorphisms also appear to have a dosage effect on AAO: the estimated mean AAO for subsets consisting of individuals carrying no ε4 allele, one ε4 allele, and two ε4 alleles, are 84.3, 75.5, and 68.8 years, respectively (Corder et al., 1993). In addition, the effect of ε4 allele on risk for LOAD is also age-dependent. The ε4 allele has a greatest impact on risk for LOAD among individuals that are between 60 and 79 years of age, while its effect is gradually attenuated as an individual reaches ages in access of 80 years (Farrer et al., 1997; Lautenschlager et al., 1999). While APOE ε4 appears to influence AD risk and AAO, it is unclear if it affects AD progression, with studies indicating it may accelerate (Craft et al., 1998; Deary et al., 2002; Dik et al., 2000; Lehtovirta et al., 1996), slow (Frisoni et al., 1995; Stern et al., 1997) or have no effect on disease progression in LOAD patients (Basun et al., 1995; Growdon et al., 1996). APOE effects have also been shown to vary widely across ethnicities (Farrer et al., 1997), and differ between males and females (Barrett, 1999; Farrer et al., 1997), even after accounting for differences in mortality between groups. APOE has also been correlated with neuropathological features of the disease: several studies report that individuals carrying two APOE ε4 alleles have higher amyloid plaque density than individuals carrying one or no APOE ε4 alleles (Hyman et al., 1996; Nagy et al., 1995; Oyama et al., 1995; Rebeck et al., 1993; Schmechel et al., 1993; Zubenko et al., 1994).

Other candidate susceptibility loci for LOAD

While APOE is a critical contributor to genetic risk, more than a third of AD cases do not carry any APOE ε4 alleles, and as AD heritability has been estimated at ~80% (Bergem, 1994), this suggests much of the heritability has not yet been characterized. At most, APOE may account for 50% of the total genetic effect in AD (Farrer et al., 1997; Roses et al., 1995; Saunders et al., 1993). To identify other genomic contributors to AD risk, a variety of approaches have been used: regional and genome-wide linkage studies (GWLS) in multiplex pedigrees, candidate gene association studies, meta-analyses of linkage and association studies, genome-wide association studies (GWAS), and most recently, whole genome sequencing (WGS) and whole exome sequencing (WES) studies.

Genome-wide linkage studies (GWLS) for LOAD have been published on multiple independent datasets since 1997 (Ashley-Koch et al., 2005; Blacker et al., 2003; Farrer et al., 2003; Hahs et al., 2006; Hiltunen et al., 2001; Kehoe et al., 1999; Mayeux, 2004; Mayeux et al., 2002; Myers et al., 2002; Pericak-Vance et al., 1997; Pericak-Vance et al., 2000). While many of these studies have focused on US and Northern European Caucasian families (Blacker et al., 2003; Hiltunen et al., 2001; Kehoe et al., 1999; Myers et al., 2002; Pericak-Vance et al., 1997; Pericak-Vance et al., 2000), other scans have looked at different ethnic groups such as Caribbean Hispanic families (Mayeux, 2004; Mayeux et al., 2002) or very large families from population isolates (Ashley-Koch et al., 2005; Farrer et al., 2003; Hahs et al., 2006), using STRs with an average spacing of 8 centiMorgans (cM). Some chromosomal regions have been studied extensively (most notably chromosomes 9, 10, and 12), although no consistently replicated AD gene have yet been identified within these regions, with the potential exception of C9ORF72 (Majounie et al., 2012). While the majority of follow-up studies have focused on these three chromosomes, several linkage regions have been identified in multiple studies that remain unexplored (Table 1, adapted from (Butler et al., 2009)).

Most genes examined as potential LOAD candidate genes have been identified due to their known biological functions, and thus have been hypothesized to play roles in AD development. Hundreds of genes have been individually tested for association with LOAD, and associations have been reported for nearly 150 genes, but few have replicated and none have gained wide acceptance (AlzGene database (http://www.alzgene.org/) accessed March 1, 2016; (Bertram et al., 2007)). Several potential causes of these inconclusive findings include small sample sizes in these studies, limited knowledge of the actual effects of genetic variation in living brains, and potential etiologic heterogeneity among the cases included in these studies.

One approach used to overcome sample size limitations is meta-analysis of independent genetic studies. A meta-analysis by Bertram and colleagues (2007) served to catalogue association results across many candidate gene association studies in the AlzGene database, as described before, but also to identify associations with AD of by systematically combining data on genetic variants that had been genotyped commonly across multiple studies. Performing meta-analyses systematically on variants genotyped in three or more samples, this study identified variants with small to modest effect sizes in twelve candidate genes for further investigation (including ACE, CHRNB2, CST3, R1, GAPDHS, IDE, MTHFR, NCSTN, PRNP, PSEN1, TF, TFAM and TNF). One of these candidates is already known to be involved in familial, early-onset forms of AD (PSEN1), suggesting that this approach may have identified valuable candidates for further examination.

Genome-wide association studies

The study by Bertram and colleagues (2007) noted a major limitation of this approach: that candidate gene studies tend to examine genes with hypothetical functional roles in disease or proximity to strong linkage signals, and this has the potential to introduce bias from several sources into meta-analytic examinations. This limitation is partially addressed by the introduction of genome-wide association studies, which deal with several potential biases by examining high-density genotyping of single nucleotide polymorphisms (SNPs) capturing most variation across the genome, reducing overrepresentation of variant data from particular genomic regions. The application of meta-analytic approaches to GWAS data represent an additional improvement upon meta-analysis in candidate genes by improving sample sizes and allowing for small to modest effects to be observed with fewer inherent potential biases, provided appropriate adjustment is made for differences between high-density genotyping platforms and for relevant issues of study design.

More than 13 studies have tested association with LOAD on high-density, genome-spanning panels of SNPs. Grupe and colleagues (2007) pooled samples and tested association with more than 17,000 gene-based putative functional SNPs across the genome, finding a signal at APOE that reached study-wide statistical significance, with multiple weaker associations observed elsewhere, many occurring in regions of known linkage. Coon and colleagues (2007) reported results of association with a half-million SNPs across the genome genotyped on over 1,000 histopathologically-verified AD cases and controls, identifying only APOE as a major susceptibility gene. A follow-up study stratifying cases by APOE genotype detected strong associations with GAB2 (MIM:606203) SNPs, and in follow-up worked observed altered GAB2 transcript levels in vulnerable neurons, and an effect of GAB2 levels on tau phosphorylation; replication studies have observed mixed results (Reiman et al., 2007). Abraham and colleagues (2008) genotyped approximately 550,000 SNPs in more than 1,000 pooled cases and 1,200 pooled controls, and in testing genetic associations, observed genome-wide statistical significance only for SNPs in or near APOE. Following up on the strongest signals that did not attain statistical significance with more individual genotyping identified the gene LRAT (MIM:604863), which is involved in the vitamin A (retinoid) cascade, a system previously implicated in AD. Bertram and colleagues (2008) analyzed 500,000 SNPs in 410 families, reporting a SNP associated with AD age-at-onset on chromosome 14q31, and providing additional evidence of associations near APOE and near GAB2. Beecham and colleagues (2009) previously reported a GWAS on approximately 550,000 SNPs in nearly 500 AD cases and 500 cognitive controls. This study confirmed genome-wide significant associations at APOE and identified a SNP on chromosome 12q13 meeting α = 5×10−5, corresponding to a false-discovery rate (FDR) < 0.20. The strongest signals with less than genome-wide statistical significance in this study were identified in regions with prior linkage evidence, suggesting that some of these are likely to underlie true associations (Avramopoulos, 2009). A GWAS of AD originating at the Mayo Clinic (Carrasquillo et al., 2009), genotyping 314,000 SNPs on 844 cases and 1,255 controls, once again verified APOE associations, and in second-stage replication analysis, identified a novel signal on the X chromosome (combined P = 3.9×10−12) in the gene PCDH11X (MIM:300246), encoding a protocadherin, a cell-cell adhesion molecule expressed in the brain. A study (Naj et al., 2010) using the original set of cases and controls from Beecham and colleagues (2009) and additional newly-identified cases and controls with high-density genotyping examined associations in a genome-wide set of markers (483,399 SNPs) identified an association with LOAD of genome-wide significance (P = 4.70×10−8) at 151.2Mb of chromosome 6q25.1 in the gene MTHFD1L, which encodes the methylenetetrahydrofolate dehydrogenase (NADP+ dependent) 1-like protein, a gene involved in the pathway synthesizing methionine from homocysteine and thus may influence homocysteine levels, known risk factors for AD. Finally, a study from the University of Pittsburgh (Kamboh et al., 2012b) with 1,291 cases and 938 controls examining ~2.5 million imputed SNPs observed its most significant novel association in the chromosome 15q21.3 gene PPP1R3B (rs3848140; P = 3.05×10−7), though this association did not attain genome-wide statistical significance. PPP1R3B is also known to be associated with high density lipoprotein and cholesterol levels.

While many of the aforementioned non-APOE associations attained genome-wide significance and were found to be association in replication datasets, associations of these variants have not been found to be statistically significant in other independent studies, and for this reason, it remains unclear whether these genetic variations are truly associated with risk. A notable limitation of these studies is the lack of statistical power to detect small or even modest associations (odds ratio (OR) < 1.5) with samples sizes of ~1,000 cases and ~1,000 controls, or less. A second-generation of GWASes with greatly increased sample sizes were able to detect the first set of associations of variants with small effects on LOAD and that replicated in multiple studies. The first of these by the European Alzheimer Disease Initiative (EADI) (Lambert et al., 2009) examined associations in a total of 6,010 LOAD cases and 8,625 controls, and observed highly significant associations in CLU (combined P = 7.5×10−9), a chromosome 8p21.1 gene which encodes clusterin or apolipoprotein J, and in CR1 (combined P = 3.7×10−9), a chromosome 1q32.2 gene encoding the complement component (3b/4b) receptor 1. Effect sizes for associations of these variants were either slightly deleterious (for CR1, OR = 1.21) or slightly protective (for CLU, OR = 0.86). A second study by the Genetic and Environmental Risk in Alzheimer Disease (GERAD) Consortium (Harold et al., 2009), released at the same time as the paper by Lambert and colleagues, also observed genome-wide statistical significance for associations of non-APOE genomic variants in their independent dataset including a total of 5,964 cases and 10,188 controls. These included strong associations in CLU (P = 8.5×10−10, OR = 0.86), one of the genes with significantly-associated variants in the study by Lambert and colleagues, and novel associations in the PICALM gene (P = 1.3×10−9, OR = 0.86). As with the previous study, the study by Harold and colleagues observed small effect sizes for variants with genome-wide significance in each of the genes (ORs ≈ 0.86), suggesting that some genetic variants contributing to LOAD other than APOE may have weak, but nevertheless real, effects on disease risk. A third study by the Cohorts for Heart and Aging Research in Genomic Epidemiology (CHARGE) consortium (Seshadri et al., 2010), which incorporated data from both the EADI and GERAD consortia to investigate an overall 9,511 cases and 28,174 controls, identified associations with genome-wide statistical significance adjacent to two sets of previously unreported loci, near BIN1 on chromosome 2q14.3 (P = 1.59×10−11, OR = 1.13) and near EXOC3L2/BLOC1S3/MARK4 on chromosome 19q13.3 (P = 6.45×10−9, OR = 1.18). BIN1 encodes the Bridging Integrator 1 protein, which is a nucleocytoplasmic adaptor protein that is heavily expressed in brain and muscle. Of the three genes proximal to the significant associations on chromosome 19q13.3 (signals which are independent of and not in linkage disequilibrium with associations in the APOE region), only two are potential biological candidates involved in AD-related pathways: BLOC1S3, which encodes subunit 3 of the biogenesis of lysosomal organelles complex-1 protein, influences endosomal to lysosomal routing, and is expressed in the brain; and MARK4, which encodes the MAP(mitogen activated protein)/microtubule affinity-regulating kinase 4 protein, is expressed exclusively in the brain, and participates in neuronal differentiation.

Meta-analysis of genome-wide association studies

The next generation of genetic studies to follow GWAS was meta-analysis across GWAS studies. By combining and meta-analyze findings within and across GWAS in large analytical consortia, and with the increase in sizes of these datasets to many tens of thousands of LOAD cases and controls, much greater power became available to identify and replicate newfound genetic associations with LOAD, and to uncover the biological mechanisms underlying disease pathology.

In 2011, the Alzheimer’s Disease Genetics Consortium (ADGC), which integrated data on 11,840 AD cases and 10,931 controls of European ancestry (Naj et al., 2011) over 15 separate genome-wide association studies of LOAD, including case-control, prospective cohort, and family-based studies, confirmed associations in APOE, PICALM, CR1, CLU, and BIN1 with genome-wide statistical significance (P<5×10−8), and identified novel associations in five other genomic regions with genome-wide statistical significance (MS4A gene cluster) or near it (P<10−6) (EPHA1, CD2AP, CD33, and ABCA7), which were validated and attained genome-wide significance with replication in silico. Similarly, in a companion paper, a collaborative GWAS effort between the three previously-mentioned consortia, EADI, GERAD, and CHARGE (Hollingworth et al., 2011), observed genome-wide significant associations at the five genes they previously identified as well as genome-wide significant association at ABCA7, and nearly significant associations at the same loci identified by the ADGC (EPHA1, CD2AP, CD33, and the MS4A gene cluster). Both groups replicated associations in each other’s consortium datasets, as well as through de novo genotyping in independent replication samples without GWAS. As both sets of analyses used genome-wide genotyping from multiple studies collected on a variety of genotyping platforms from both Illumina and Affymetrix, all GWAS datasets were imputed using haplotype reference data from the International HapMap Project (International HapMap et al., 2007) to examine a common set of approximately 2.5 million SNPs across the genome. A number of subsequent studies provided independent replications of these findings (Biffi et al., 2010; Carrasquillo et al., 2010; Carrasquillo et al., 2011a; Carrasquillo et al., 2011b; Corneveaux et al., 2010; Gu et al., 2011; Jun et al., 2010; Kamboh et al., 2012c; Lambert et al., 2011; Omoumi et al., 2014; Piaceri et al., 2011).

To maximize sample size and thereby power to detect LOAD susceptibility loci with even more modest effects, the four consortia, ADGC, EADI, GERAD, and CHARGE, combined data and analytical resources to form a mega-consortium, the International Genomics of Alzheimer’s Project (IGAP), to conduct a new GWAS. This GWAS was performed on 17,008 AD cases and 37,154 controls of European ancestry, with de novo genotyping of variants with genome-wide significant and suggestive associations (P<0.001) in a large replication dataset of 8,572 AD cases and 11,312 cognitive controls from multiple European datasets without available GWAS (Lambert et al., 2013b). Except for CD33, all previously-identified GWAS loci reached genome-wide significance in the discovery analysis, and an additional 11 loci (HLA-DRB5/DRB1, PTK2B, SORL1, SLC24A4-RIN3, INPP5D, MEF2C, NME8, ZCWPW1, CELF1, FERMT2, and CASS4) demonstrated genome-wide statistically significant associations with the inclusion of analysis of replication datasets. Among these, SORL1 had been identified in a GWAS of Japanese, Korean, and Caucasian AD cases and controls (the lattermost taken from the ADGC), however, as a well-established functional candidate, it had not been observed in a GWAS of European ancestry until the IGAP analyses, likely due to its low minor allele frequency (MAF) of ~0.04. Subsequent replication efforts have provided confirmation of the signal in ZCWPW1 (Ruiz et al., 2014) as well as evidence of association at TRIP4. Multiple validation studies (studies examining associations among different ethnicities) in Southeast Asian populations have demonstrated associations at many of these loci: CLU (Komatsu et al., 2011; Liu et al., 2014; Yu et al., 2013), CR1 (Jin et al., 2012), PICALM (Chung et al., 2013; Liu et al., 2013; Miyashita et al., 2013), BIN1 (Liu et al., 2013; Miyashita et al., 2013), and the MS4A gene cluster (Deng et al., 2012; Tan et al., 2013).

In addition to the Japanese/Korean/Caucasian meta-analysis previously described, multiple GWAS meta-analysis have been performed in non-Caucasian ethnicities, though sample sizes of these studies have been considerably smaller than those in the previously described studies. A GWAS of AD in 544 affected and 549 unaffected samples of Caribbean Hispanic ancestry (Tan et al., 2013) identified a number of strong but not genome-wide statistically significant novel associations (P < 9×10−6) and replicated associations at CLU, PICALM, and BIN1. A GWAS of AD in 5,896 African American subjects (including 1,968 cases and 3,928 controls) reported genome-wide significant associations in ABCA7 (Reitz et al., 2013), with similarly strong signals in two adjacent loci, HMHA1 and GRIN3B (likely part of the same signal), around 1-1.05Mb on chromosome 19p13.3. Gene-based analyses identified nominal associations of CR1, BIN1, EPHA1, and CD33.

Genome-wide association studies and meta-analyses of AD-related phenotypes

Genetic studies of phenotypes related to AD, including endophenotypes (phenotypes endogenous to the disease pathway (Flint and Munafo, 2007; Rice et al., 2001), are among alternative approaches to identifying genetic risk factors for AD. Alternative phenotypes for AD include characteristics of disease (e.g., age-at-onset), biomarkers capturing or correlating with disease pathology (e.g., cerebrospinal fluid (CSF) Aβ42), and modified case definitions (e.g., cases with onset after 85 years of age). Since the advent of whole genome studies, a multitude of studies examining AD-related phenotypes have used genome-wide association in efforts to better characterize genomic contributors to AD.

Among the most thoroughly investigated endophenotypes are CSF Aβ42 and CSF tau protein levels. An early CSF GWAS (Han et al., 2010) in the Alzheimer’s Disease Neuroimaging Intitiative (ADNI) dataset did not find any genome-wide statistically significant associations with CSF biomarkers Aβ1-42, T-tau, and P-tau181P among AD cases, however two associations were observed to meet that threshold in cognitively normal individuals, one between a variant in CYP19A1 (rs2899472, P = 1.86×10−9) and Aβ1-42, and another between a chromosome 12 SNP upstream of SRRM4 (rs1997111, P = 1.10×10−8) and T-tau. CYP1A1 encodes aromatase, which is involved in estrogen regulation, while SRRM4 regulates alternative splicing of microexons in the brain, and misregulation of which has been linked to autism spectrum disorders (Irimia et al., 2014). A subsequent reanalysis of these same data (Kim et al., 2011) identified genome-wide significant association with T-tau in the chromosome 2 gene EPC2 (rs4499362, P ≈ 2.40×10−8), which has been implicated in syndromes leading to mental retardation (van Bon et al., 2010). More recently, a GWAS of CSF biomarkers (Cruchaga et al., 2013) identified four genome-wide significant associations associated with CSF tau levels (including Tau and pTau): APOE; a chromosome 3q28 SNP rs9877502 (P = 4.89×10−9 for Tau; located between GEMC1 and OSTN); a chromosome 9p24.2 SNP rs514716 located at 9p24.2 in the gene GLIS3 (P = 1.07×10−8 for Tau and P = 3.22×10−9 for pTau); and a chromosome 6p21.1 SNP rs6922617 (P = 3.58×10−8 for CSF pTau) within the TREM gene cluster. Follow-up analyses on these SNPs showed rs9877502 to also be associated with AD (P = 2.67×10−4) and cognitive decline (P = 4.86×10−5), suggesting the identification of a novel, if weak, contributor to AD risk.

Among the more widely collected and more frequently studied AD-related phenotypes is age-at-onset (AAO) of the disease, which is a marker for both risk and severity of AD. A GWAS of AAO by Kamboh and colleagues (2012a) of 2,222 AD cases identified no genome-wide significant associations outside of the APOE region, with the strongest non-APOE association being observed on chromosome 4q31.3 in the gene DCHS2 (rs1466662; P = 4.95×10−7). A subsequent GWAS meta-analysis of 9,162 subjects (Naj et al., 2014), which included the Kamboh dataset, identified genome-wide significant associations only in and around the APOE locus, with the next most significant association in the chromosome 13q33.3 gene MYO16 (rs9521011; P = 7.62×10−8). This study examined the overall contribution of the most significantly associated variants identified in the ADGC GWAS to variation in AAO, and determined that while APOE contributes to 3.9% of variation in AAO, the combined effect of all other risk loci contribute to 1.1% of AAO variation, suggesting that no AD risk loci modulate AAO as strongly as APOE.

Multiple GWAS of brain imaging datasets have been undertaken to characterize pathological progression in living subjects. An GWAS on hippocampal atrophy performed using ADNI data (Potkin et al., 2009) identified genome-wide significant associations only in the APOE region, with several candidate genes demonstrating associations in the range of P < 10−6: EFNA5, CAND1, MAGI2, ARSB, and PRUNE2. A GWAS in the MIRAGE study (Melville et al., 2012) examined phenotypes related to hippocampal atrophy, in particular hippocampal volume (HV), total cerebral volume (TCV), and white matter hyperintensities (WMH), in 981 Caucasian and 419 African-American AD cases matched with cognitively normal siblings, with replication in the ADNI cohort. Novel genome-wide significant associations (P < 5×10−8) were attained for HV with SNPs in the APOE, F5/SELP, LHFP and GCFC2 gene regions. Associations with different SNPs in the same gene (P < 10−5 in Caucasians and P < 2.2×10−4 in African Americans) were also observed for PICALM (now a confirmed AD risk locus) with HV, SYNPR with TCV and TTC27 with WMH. A 2010 GWAS (Shen et al., 2010) was performed on 142 measures of grey matter (GM) density, volume, and cortical thickness taken on 733 ADNI participants (175 AD cases, 354 individuals with amnestic mild cognitive impairment (MCI), and 204 congitively normal controls). Using hierarchical clustering and heat maps to analyze the results, investigators identified association in APOE with multiple brain regions, variants in EPHA4 (rs10932886) showed nominal associations (P < 10−6) with 27 voxel-based morphometry (VBM) phenotypes, while variants in TP63 (rs7610017) and NXPH1 (rs6463843) demonstrated associations with 19 and nine VBM phenotypes, respectively. EPHA4 belongs to the same subfamily of ephrin receptors as the confirmed GWAS locus, EPHA1, whereas NXPH1 encodes a protein which protein forms a very tight complex with alpha neurexins, a group of proteins that promote adhesion between dendrites and axons (Missler and Sudhof, 1998). A voxelwise genome-wide association study (vGWAS) (Stein et al., 2010) on the same data investigating morphometry differences in each of 31,622 voxels of the entire brain across 740 elderly subjects with AD, MCI, or normal cognition, did not identify any novel genome-wide associations in large part due to stringent multiple hypothesis testing correction. The investigators did find a common variant in the GRIN2B glutamate receptor gene to be over-represented in Alzheimer’s disease and was associated with ∼1.5% lower temporal lobe volume per risk allele in the elderly (P < 5×10−7). Most recently, a large GWAS of HVs on >20,000 cognitively-normal individuals (Bis et al., 2012) identified several possible AD candidate genes with effects on HV, which are involved in a number of processes: HRK (associated with apoptosis), LEMD3 (transforming growth factor antagonism), ASTN2 (neuronal migration), MSRB3 (oxidative stress), WIF1 (brain development), FBXW8 (ubiquitin pathway), and DPP4 (incretin pathway).

Haplotype, gene-based, and pathway analyses

Single variant analyses have been typically used to identify AD risk loci, but there may be limitations to examining only singe variants in each analysis. Among the strongest of these is the considerable correction for Type I error required (and thus larger sample size required) in order to account for multiple hypothesis testing. Other limitations include that if susceptibility is conferred by multiple different variants within a locus (Ioannidis, 2007; Neale and Sham, 2004), complex patterns of association might be observed where the same variants have different associations across different datasets despite similar expectations from power estimations (Moskvina and O’Donovan, 2007). Yet another limitation is that ungenotyped yet causal rare variants may not be captured adequately through ‘tagging’ approaches utilizing linkage disequilibrium. One way to address this last limitation is to phase alleles at multiple adjacent variants into haplotypes, especially if genotyping or sequencing rare variants is too costly for the study, or if there is evidence that co-transmission of alleles at different loci on the same chromosomal arm may uniquely elevate risk for disease. One such study to perform association analyses of haplotypes in AD utilized data from the GERAD and EADI consortia (Lambert et al., 2013a), with discovery analyses in EADI GWAS data (2,025 AD cases/5,328 controls), replication in GERAD GWAS data (2,820 AD cases/6,356 controls), and further replication by custom genotyping in five additional case-control datasets (including 5,093 AD cases/4,061 controls). This study reported genome-wide significant association at a haplotype in the chromosome 10p13 gene FRMD4A (OR (95% CI): 1.68 (1.43, 1.96); P = 1.1×10−10), for which secondary analyses demonstrated that FRMD4A polymorphisms were associated with plasma Aβ42/Aβ40 ratio (best signal, P = 5.4×10−7). Subsequent work found that the down-regulation of FRMD4A is associated with increased APP-β-secretase interaction, increased Aβ40 secretion, and altered phosphorylation of tau (Martiskainen et al., 2015).

In situations where AD risk is driven by multiple different variants within a gene but there is inadequate power to detect individual effects in single variant analyses, an alternative approach for examining association is by performing gene-wide tests, or multilocus association tests that include all variants in the exonic and intronic regions of a genes, as well as potentially regulatory intergenic variants upstream or downstream of the gene in the immediate vicinity. Genome-wide gene-based analyses like those using imputed common variant data have also been performed in AD using the IGAP consortium data to identify genes with multiple strong variant associations that failed to meet genome-wide statistical significance when examined separately (Escott-Price et al., 2014). These analyses identified five genes where gene-wide association are P < 2.56×10−6 but where SNPs in single-variant analyses were not associated at P < 5×10−8, including TP53INP1 (gene-wide P = 1.4×10−6) and IGHV1-67 (gene-wide P = 7.9×10−8), and three genes near loci with genome-wide significant association signals in single variant analysis (ZNF3, gene-wide P = 8.6×10−7; NDUFS3, gene-wide P = 4.8×10−7; MTCH2, gene-wide P = 2.5×10−6). Evidence exists for a role in neurodegeneration for at least one of these genes, NDUFS3, for which a deficiency is known to cause Leigh syndrome (Benit et al., 2004), an early-onset, progressive neurodegenerative disorder with a neuropathology featuring focal lesions including areas of gliosis.

Beyond the identification of groups of variants or whole genes contributing to elevated AD susceptibility, the ultimate goal of these genomic studies are to identify the key biological pathways influencing development of AD as targets for the development of therapeutic interventions to treat and ideally cure the disease. Yet another study design attempts to ask this question directly: using existing knowledge about biological pathways and which genes are involved in these pathways, is it possible to identify the most important pathways to target for developing treatments? Pathway analyses attempt this by combining associations across multiple variants from GWAS data not at the gene-level, but at the pathway-level by examining variants in all genes clustered in defined biological pathways. Pathway gene-sets can be derived from a variety of sources, including using network analysis tools to identify gene-sets from pathologically-relevant data (e.g., co-expression networks in disease-specific tissues), or curated in publicly-available databases such as the KEGG (Kyoto Encyclopedia of Genes and Genomes) and GO (Gene Ontology) resources. Among the early pathway analyses performed, a 2010 study by Jones and colleagues (2010) utilized GERAD GWAS data (in which the first genome-wide associations of CLU, CR1, and PICALM were observed) and a gene-set enrichment analysis (GSEA) approach as implemented in the program ALLIGATOR (Wang et al., 2007) to identify gene-sets with stronger associations than expected by chance alone. This approach found a significant overrepresentation of association signals in pathways related to cholesterol metabolism (strongest pathway: cholesterol transport [GO:0030301], P < 10−5) and immune response (strongest pathway: complement activation [GO:0006958], P = 2×10−5). An analysis in the ADNI GWAS dataset (Ramanan et al., 2012) used a pathway enrichment analysis based on a composite measure of memory generated applying modern psychometric theory to item-level data from the ADNI neuropsychological test battery, and identified 27 canonical pathways that were enriched (FDR-corrected P < 0.05) against this memory score, including cell adhesion, neuronal differentiation, and inflammation-related signaling pathways. More recently, a network analysis by Zhang and colleagues (2013) examined gene-regulatory networks in 1,647 postmortem brain tissues from LOAD cases and unaffected individuals, and demonstrated reconfiguration of the molecular interaction structure in specific ways in LOAD, which highlighted genes involved in pathogen phagocytosis regulated by TYROBP, which is upregulated in AD. Expression of TYROBP is restricted to cells involved in innate immunity, including microglia in the brain (Schleinitz et al., 2009).

Rare variant association studies and the future of AD genetic studies

Multilocus association studies like those described in the previous section have the potential to capture some of the contribution from low-frequency (0.005 ≤ MAF < 0.05) and rare (MAF < 0.005) genomic variants, however capturing genotypes on rare variants in large datasets have until the last 5 years remained a prohibitive challenge, either due to technological limitations or cost. In addition to cost-lowering advances in WGS and WES, technologies such as the exome array have greatly improved the ability to capture genotypes on hundreds of thousands to millions of rare variants throughout the genome in large datasets. While these approaches have advanced the ability to characterize rare variation in study participants, there still remain substantial analytical limitations, most of these stemming from the lack of adequate sample sizes to detect associations of variants with MAF < 0.01 at genome-wide statistical significance levels.

Thus far, exome array studies of AD have lacked adequate sample sizes to detect novel low-frequency and rare variant associations (Chen et al., 2015; Chung et al., 2014), despite the relatively small number of coding region variants genotyped (~250,000) on these platforms. A large-scale exome array study in the IGAP consortium is nearing completion with discovery data on ~17,000 AD cases and cognitive controls and may identify novel rare variants using this approach.

Exome and genome sequencing approaches have recently yielded novel insights into the genetic contributors to AD. Rare loss-of-function mutations R47H and R62H in the chromosome 6p21.1 gene TREM2 have recently been found to lead to an increase of as much as 400% in AD risk (Guerreiro et al., 2013b; Jonsson et al., 2013). TREM2 implicates innate immunity pathways (Golde et al., 2013), is involved in the regulation of phagocytic pathways, and has been found to inhibit cytokine production and secretion in microglia, reducing inflammatory response (Jiang et al., 2013). Mutations in TREM2 have been previously implicated in Nasu Hakola disease (Paloneva et al., 2002) and early-onset dementia without bone cysts (Chouery et al., 2008). The identification of this gene has reinforced the growing consensus that “neuroinflammation” as driven by microglia is a critical component of AD pathogenesis. A whole-genome sequencing study examining genotypes of 1,795 Icelanders found a coding mutation in the APP gene (A673T) that protects against Alzheimer’s disease and cognitive decline in the elderly without Alzheimer’s disease (Jonsson et al., 2012). SORL1, which was first an AD candidate gene (Rogaeva et al., 2007) and then observed in the IGAP GWAS of AD, was recently found through exome sequencing to carry mutations causing a form of early-onset AD with autosomal dominant inheritance (Pottier et al., 2012). Exome sequencing of a Turkish family with AD identified a mutation in NOTCH3 (R1231C) (Guerreiro et al., 2012), which has previously been implicated in a subtype of vascular dementia, CADASIL (cerebral autosomal dominant arteriopathy with subcortical infarcts and leukoencephalopathy). A 2014 study by Cruchaga and colleagues (2014) identified rare coding variants in the chromosome 19q13.2 gene PLD3, encoding an enzyme of the phospholipase D family, however this association not been observed consistently in follow-up studies (Cruchaga and Goate, 2015; Heilmann et al., 2015; Hooli et al., 2015; Lambert et al., 2015; van der Lee et al., 2015).

The next generation of rare variant studies in AD is currently in development and includes the integration of multiple analytic strategies for the identification of novel variants affecting Alzheimer’s disease risk, including linkage and association. The upcoming Alzheimer’s Disease Sequencing Project (ADSP; https://www.niagads.org/adsp/content/home) is a joint project by the National Institute on Aging and National Human Genome Research Institute to identify novel risk and protective variants for AD in nearly 600 whole genomes of from 110 multiplex pedigrees heavily burdened with AD and approximately 10,500 whole exomes on unrelated AD cases and non-demented controls. Preliminary work on the sample sets being incorporated into the ADSP includes whole genome linkage analyses using GWAS data on multiplex White and Caribbean Hispanic pedigrees to identify candidate genomic regions for particular focus in an upcoming genome scan of WGS data (Barral et al., 2015; Kunkle et al., 2016). These and other studies promise to do much to characterize the yet-undetected heritable components in Alzheimer’s susceptibility, the “dark matter” of the genomic risk of AD.

Table 2.

Genomic risk loci for sporadic/late-onset Alzheimer’s disease observed in linkage and/or association studies (adapted from Chouraki and Seshadri (2014) and Guerreiro and colleagues (2013a))

Locus Variant Chr Position (bp) (hg19) Effect/Alt Alleles Effect Allele Freq OR (95% CI) Potential Biological Process or Pathways Affected References
Common risk loci identified through linkage
APOE rs429358/C112R 19 45411941 ε4/(ε3/ε2) 0.14 ∼2 to 5 Cholesterol/lipid metabolism (Corder et al., 1993; Saunders et al., 1993; Strittmatter et al., 1993)
rs7412/C158R 19 45412079
Common risk loci identified through GWAS
CR1 rs6656401 1 207692049 A/G 0.20 1.18 (1.14–1.22) Immune and complement systems/inflammatory response (Lambert et al., 2009)
BIN1 rs6733839 2 127892810 T/C 0.41 1.22 (1.18–1.25) Synaptic vesicle endocytosis (Seshadri et al., 2010)
INPP5D rs35349669 2 234068476 T/C 0.49 1.08 (1.05–1.11) Immune response and inflammation; APP metabolism (Lambert et al., 2013b)
MEF2C rs190982 5 88223420 G/A 0.41 0.93 (0.90–0.95) Immune response and inflammation; Hippocampal synaptic function (Lambert et al., 2013b)
HLA-DRB1/5 rs9271192 6 32578530 C/A 0.28 1.11 (1.08–1.15) Immune response and inflammation (Lambert et al., 2013b)
CD2AP rs10948363 6 47487762 G/A 0.27 1.10 (1.07–1.13) Receptor endocytosis and cytokinesis (Hollingworth et al., 2011; Naj et al., 2011)
NME8 rs2718058 7 37841534 G/A 0.37 0.93 (0.90–0.95) Cytoskeletal function and axonal transport (Lambert et al., 2013b)
ZCWPW1 rs1476679 7 100004446 C/T 0.29 0.91 (0.89–0.94) Epigenetic regulation (Lambert et al., 2013b)
EPHA1 rs11771145 7 143110762 A/G 0.34 0.90 (0.88–0.93) Immune and complement systems/inflammatory response (Hollingworth et al., 2011; Naj et al., 2011)
PTK2B rs28834970 8 27195121 C/T 0.37 1.10 (1.08–1.13) Hippocampal synaptic function; cell migration (Lambert et al., 2013b)
CLU rs9331896 8 27467686 C/T 0.38 0.86 (0.84-0.89) Immune and complement systems/inflammatory response; cholesterol/lipid metabolism (Harold et al., 2009; Lambert et al., 2009)
CELF1 rs10838725 11 47557871 C/T 0.32 1.08 (1.05–1.11) Cytoskeletal function and axonal transport (Lambert et al., 2013b)
MS4A (gene cluster) rs983392 11 59923508 G/A 0.40 0.90 (0.87-0.92) Immune and complement systems/inflammatory response (Hollingworth et al., 2011; Naj et al., 2011)
PICALM rs10792832 11 85867875 A/G 0.36 0.87 (0.85–0.89) Clathrin-mediated endocytosis (Harold et al., 2009)
SORL1 rs11218343 11 121435587 C/T 0.039 0.77 (0.72–0.82) Endocytosis and cargo sorting (trafficking and metabolism of APP), lipid transport (Lambert et al., 2013b)
FERMT2 rs17125944 14 53400629 C/T 0.092 1.14 (1.09–1.19) Angiogenesis, Tau pathology (Lambert et al., 2013b)
SLC24A4- RIN3 rs10498633 14 92926952 T/G 0.22 0.91 (0.88–0.94) Possible cardiovascular link (Lambert et al., 2013b)
ABCA7 rs4147929 19 1063443 A/G 0.19 1.15 (1.11–1.19) Cholesterol/lipid metabolism; Immune and complement systems/inflammatory response (Hollingworth et al., 2011; Naj et al., 2011)
CASS4 rs7274581 20 55018260 C/T 0.083 0.88 (0.84–0.92) APP and Tau pathology; Cytoskeletal function and axonal transport (Lambert et al., 2013b)
Rare Variant Risk Loci
APP rs63750847/A673T 21 27269932 A/G 0.0045 21 Neuronal Differentiation/Amyloid cascade (Jonsson et al., 2012)
TREM2 rs75932628/R47H 6 41129252 T/C 0.0063 2.26 (1.71–2.98) Immune response; chronic inflammation (Guerreiro et al., 2013b; Jonsson et al., 2013)
Risk loci from GWAS independently replicated
TREML2 rs9381040 6 41154650 C/T 0.30. 0.93 (0.91–0.96) Inflammation regulation (Benitez et al., 2013)
TRIP4 rs74615166 15 64725490 C/T 0.02 1.31 (1.17–1.42) Immune response; endocrine processes (Ruiz et al., 2014)
Potential AD loci from GWAS not yet replicated
ACE rs138190086 17 61538148 A/G 0.011 1.34 (1.20–1.50) Regulates blood pressure; part of the renin-angiotensins system (Lambert et al., 2013b)
Potential AD loci with inconsistent associations
CD33 rs3865444 19 51727962 A/C 0.31 0.94 (0.91–0.96) Immune and complement systems/inflammatory response (Hollingworth et al., 2011; Naj et al., 2011)
DSG2 rs8093731 18 29088958 C/T 0.017 0.73 (0.62–0.86) Mediates cell-cell junctions between multiple cell types (Lambert et al., 2013b)
PLD3 rs145999145 19 40877595 A/G 0.060 2.75 (2.05–3.68) Hydrolysis of membrane phospholipids; APP processing (Cruchaga et al., 2014)

Acknowledgments

This work is supported through a grant U01 AG032984 from the National Institutes of Health, National Institute on Aging (NIH-NIA).

Footnotes

The authors have no conflicts of interest to declare.

References

  1. Abraham R, Moskvina V, Sims R, Hollingworth P, Morgan A, Georgieva L, Dowzell K, Cichon S, Hillmer AM, O’Donovan MC, et al. A genome-wide association study for late-onset Alzheimer’s disease using DNA pooling. BMC Med Genomics. 2008;1:44. doi: 10.1186/1755-8794-1-44. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Alzheimer A. Über eine eigenartige Erkrankung der Hirnrinde. Allgemeine Zeitschrift für Psychiatrie und psychisch-gerichtliche Medizin. 1907;64 [Google Scholar]
  3. Alzheimer’s Association. 2015 Alzheimer’s Disease Facts and Figures. Washington, D.C.: 2015. [DOI] [PubMed] [Google Scholar]
  4. Ashley-Koch AE, Shao Y, Rimmler JB, Gaskell PC, Welsh-Bohmer KA, Jackson CE, Scott WK, Haines JL, Pericak-Vance MA. An autosomal genomic screen for dementia in an extended Amish family. Neurosci Lett. 2005;379(3):199–204. doi: 10.1016/j.neulet.2004.12.065. [DOI] [PubMed] [Google Scholar]
  5. Avramopoulos D. Genetics of Alzheimer’s disease: recent advances. Genome Med. 2009;1(3):34. doi: 10.1186/gm34. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Bachman DL, Wolf PA, Linn R, Knoefel JE, Cobb J, Belanger A, D’Agostino RB, White LR. Prevalence of dementia and probable senile dementia of the Alzheimer type in the Framingham Study. Neurology. 1992;42(1):115–9. doi: 10.1212/wnl.42.1.115. [DOI] [PubMed] [Google Scholar]
  7. Barral S, Cheng R, Reitz C, Vardarajan B, Lee J, Kunkle B, Beecham G, Cantwell LS, Pericak-Vance MA, Farrer LA, et al. Linkage analyses in Caribbean Hispanic families identify novel loci associated with familial late-onset Alzheimer’s disease. Alzheimers Dement. 2015;11(12):1397–406. doi: 10.1016/j.jalz.2015.07.487. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Barrett AM. Probable Alzheimer’s disease: gender-related issues. J Gend Specif Med. 1999;2(1):55–60. [PubMed] [Google Scholar]
  9. Basun H, Grut M, Winblad B, Lannfelt L. Apolipoprotein epsilon 4 allele and disease progression in patients with late-onset Alzheimer’s disease. Neurosci Lett. 1995;183(1–2):32–4. doi: 10.1016/0304-3940(94)11107-t. [DOI] [PubMed] [Google Scholar]
  10. Beecham GW, Martin ER, Li YJ, Slifer MA, Gilbert JR, Haines JL, Pericak-Vance MA. Genome-wide association study implicates a chromosome 12 risk locus for late-onset Alzheimer disease. Am J Hum Genet. 2009;84(1):35–43. doi: 10.1016/j.ajhg.2008.12.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Benit P, Slama A, Cartault F, Giurgea I, Chretien D, Lebon S, Marsac C, Munnich A, Rotig A, Rustin P. Mutant NDUFS3 subunit of mitochondrial complex I causes Leigh syndrome. J Med Genet. 2004;41(1):14–7. doi: 10.1136/jmg.2003.014316. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Benitez BA, Jin SC, Guerreiro R, Graham R, Lord J, Harold D, Sims R, Lambert JC, Gibbs JR, Bras J, et al. Missense variant in TREML2 protects against Alzheimer’s disease. Neurobiol Aging. 2013 doi: 10.1016/j.neurobiolaging.2013.12.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Bergem AL. Heredity in dementia of the Alzheimer type. Clin Genet. 1994;46(1 Spec):144–9. doi: 10.1111/j.1399-0004.1994.tb04216.x. [DOI] [PubMed] [Google Scholar]
  14. Bergem AL, Engedal K, Kringlen E. The role of heredity in late-onset Alzheimer disease and vascular dementia. A twin study. Arch Gen Psychiatry. 1997;54(3):264–70. doi: 10.1001/archpsyc.1997.01830150090013. [DOI] [PubMed] [Google Scholar]
  15. Bertram L, Lange C, Mullin K, Parkinson M, Hsiao M, Hogan MF, Schjeide BM, Hooli B, Divito J, Ionita I, et al. Genome-wide association analysis reveals putative Alzheimer’s disease susceptibility loci in addition to APOE. Am J Hum Genet. 2008;83(5):623–32. doi: 10.1016/j.ajhg.2008.10.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Bertram L, McQueen MB, Mullin K, Blacker D, Tanzi RE. Systematic meta-analyses of Alzheimer disease genetic association studies: the AlzGene database. Nat Genet. 2007;39(1):17–23. doi: 10.1038/ng1934. [DOI] [PubMed] [Google Scholar]
  17. Biffi A, Anderson CD, Desikan RS, Sabuncu M, Cortellini L, Schmansky N, Salat D, Rosand J, Alzheimer’s Disease Neuroimaging I Genetic variation and neuroimaging measures in Alzheimer disease. Arch Neurol. 2010;67(6):677–85. doi: 10.1001/archneurol.2010.108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Bird TD, Lampe TH, Nemens EJ, Miner GW, Sumi SM, Schellenberg GD. Familial Alzheimer’s disease in American descendants of the Volga Germans: probable genetic founder effect. Ann Neurol. 1988;23(1):25–31. doi: 10.1002/ana.410230106. [DOI] [PubMed] [Google Scholar]
  19. Bis JC, DeCarli C, Smith AV, van der Lijn F, Crivello F, Fornage M, Debette S, Shulman JM, Schmidt H, Srikanth V, et al. Common variants at 12q14 and 12q24 are associated with hippocampal volume. Nat Genet. 2012;44(5):545–51. doi: 10.1038/ng.2237. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Blacker D, Bertram L, Saunders AJ, Moscarillo TJ, Albert MS, Wiener H, Perry RT, Collins JS, Harrell LE, Go RC, et al. Results of a high-resolution genome screen of 437 Alzheimer’s disease families. Hum Mol Genet. 2003;12(1):23–32. doi: 10.1093/hmg/ddg007. [DOI] [PubMed] [Google Scholar]
  21. Blessed G, Tomlinson BE, Roth M. The association between quantitative measures of dementia and of senile change in the cerebral grey matter of elderly subjects. Br J Psychiatry. 1968;114(512):797–811. doi: 10.1192/bjp.114.512.797. [DOI] [PubMed] [Google Scholar]
  22. Breitner JC, Gatz M, Bergem AL, Christian JC, Mortimer JA, McClearn GE, Heston LL, Welsh KA, Anthony JC, Folstein MF, et al. Use of twin cohorts for research in Alzheimer’s disease. Neurology. 1993;43(2):261–7. doi: 10.1212/wnl.43.2.261. [DOI] [PubMed] [Google Scholar]
  23. Breitner JC, Welsh KA, Gau BA, McDonald WM, Steffens DC, Saunders AM, Magruder KM, Helms MJ, Plassman BL, Folstein MF, et al. Alzheimer’s disease in the National Academy of Sciences-National Research Council Registry of Aging Twin Veterans. III. Detection of cases, longitudinal results, and observations on twin concordance. Arch Neurol. 1995;52(8):763–71. doi: 10.1001/archneur.1995.00540320035011. [DOI] [PubMed] [Google Scholar]
  24. Butler AW, Ng MY, Hamshere ML, Forabosco P, Wroe R, Al-Chalabi A, Lewis CM, Powell JF. Meta-analysis of linkage studies for Alzheimer’s disease–a web resource. Neurobiol Aging. 2009;30(7):1037–47. doi: 10.1016/j.neurobiolaging.2009.03.013. [DOI] [PubMed] [Google Scholar]
  25. Carrasquillo MM, Belbin O, Hunter TA, Ma L, Bisceglio GD, Zou F, Crook JE, Pankratz VS, Dickson DW, Graff-Radford NR, et al. Replication of CLU, CR1, and PICALM associations with alzheimer disease. Arch Neurol. 2010;67(8):961–4. doi: 10.1001/archneurol.2010.147. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Carrasquillo MM, Belbin O, Hunter TA, Ma L, Bisceglio GD, Zou F, Crook JE, Pankratz VS, Sando SB, Aasly JO, et al. Replication of BIN1 association with Alzheimer’s disease and evaluation of genetic interactions. J Alzheimers Dis. 2011a;24(4):751–8. doi: 10.3233/JAD-2011-101932. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Carrasquillo MM, Belbin O, Hunter TA, Ma L, Bisceglio GD, Zou F, Crook JE, Pankratz VS, Sando SB, Aasly JO, et al. Replication of EPHA1 and CD33 associations with late-onset Alzheimer’s disease: a multi-centre case-control study. Mol Neurodegener. 2011b;6(1):54. doi: 10.1186/1750-1326-6-54. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Carrasquillo MM, Zou F, Pankratz VS, Wilcox SL, Ma L, Walker LP, Younkin SG, Younkin CS, Younkin LH, Bisceglio GD, et al. Genetic variation in PCDH11X is associated with susceptibility to late-onset Alzheimer’s disease. Nat Genet. 2009;41(2):192–8. doi: 10.1038/ng.305. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Chen JA, Wang Q, Davis-Turak J, Li Y, Karydas AM, Hsu SC, Sears RL, Chatzopoulou D, Huang AY, Wojta KJ, et al. A multiancestral genome-wide exome array study of Alzheimer disease, frontotemporal dementia, and progressive supranuclear palsy. JAMA Neurol. 2015;72(4):414–22. doi: 10.1001/jamaneurol.2014.4040. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Chouery E, Delague V, Bergougnoux A, Koussa S, Serre JL, Megarbane A. Mutations in TREM2 lead to pure early-onset dementia without bone cysts. Hum Mutat. 2008;29(9):E194–204. doi: 10.1002/humu.20836. [DOI] [PubMed] [Google Scholar]
  31. Chouraki V, Seshadri S. Genetics of Alzheimer’s disease. Adv Genet. 2014;87:245–94. doi: 10.1016/B978-0-12-800149-3.00005-6. [DOI] [PubMed] [Google Scholar]
  32. Chung SJ, Kim MJ, Kim J, Kim YJ, You S, Koh J, Kim SY, Lee JH. Exome array study did not identify novel variants in Alzheimer’s disease. Neurobiol Aging. 2014;35(8):1958 e13–4. doi: 10.1016/j.neurobiolaging.2014.03.007. [DOI] [PubMed] [Google Scholar]
  33. Chung SJ, Lee JH, Kim SY, You S, Kim MJ, Lee JY, Koh J. Association of GWAS top hits with late-onset Alzheimer disease in Korean population. Alzheimer Dis Assoc Disord. 2013;27(3):250–7. doi: 10.1097/WAD.0b013e31826d7281. [DOI] [PubMed] [Google Scholar]
  34. Coon KD, Myers AJ, Craig DW, Webster JA, Pearson JV, Lince DH, Zismann VL, Beach TG, Leung D, Bryden L, et al. A high-density whole-genome association study reveals that APOE is the major susceptibility gene for sporadic late-onset Alzheimer’s disease. J Clin Psychiatry. 2007;68(4):613–8. doi: 10.4088/jcp.v68n0419. [DOI] [PubMed] [Google Scholar]
  35. Corder EH, Saunders AM, Risch NJ, Strittmatter WJ, Schmechel DE, Gaskell PC, Jr, Rimmler JB, Locke PA, Conneally PM, Schmader KE, et al. Protective effect of apolipoprotein E type 2 allele for late onset Alzheimer disease. Nat Genet. 1994;7(2):180–4. doi: 10.1038/ng0694-180. [DOI] [PubMed] [Google Scholar]
  36. Corder EH, Saunders AM, Strittmatter WJ, Schmechel DE, Gaskell PC, Small GW, Roses AD, Haines JL, Pericak-Vance MA. Gene dose of apolipoprotein E type 4 allele and the risk of Alzheimer’s disease in late onset families. Science. 1993;261(5123):921–3. doi: 10.1126/science.8346443. [DOI] [PubMed] [Google Scholar]
  37. Corneveaux JJ, Myers AJ, Allen AN, Pruzin JJ, Ramirez M, Engel A, Nalls MA, Chen K, Lee W, Chewning K, et al. Association of CR1, CLU and PICALM with Alzheimer’s disease in a cohort of clinically characterized and neuropathologically verified individuals. Hum Mol Genet. 2010;19(16):3295–301. doi: 10.1093/hmg/ddq221. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Craft S, Teri L, Edland SD, Kukull WA, Schellenberg G, McCormick WC, Bowen JD, Larson EB. Accelerated decline in apolipoprotein E-epsilon4 homozygotes with Alzheimer’s disease. Neurology. 1998;51(1):149–53. doi: 10.1212/wnl.51.1.149. [DOI] [PubMed] [Google Scholar]
  39. Cruchaga C, Goate AM. Cruchaga & Goate reply. Nature. 2015;520(7545):E5–6. doi: 10.1038/nature14037. [DOI] [PubMed] [Google Scholar]
  40. Cruchaga C, Karch CM, Jin SC, Benitez BA, Cai Y, Guerreiro R, Harari O, Norton J, Budde J, Bertelsen S, et al. Rare coding variants in the phospholipase D3 gene confer risk for Alzheimer’s disease. Nature. 2014;505(7484):550–4. doi: 10.1038/nature12825. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Cruchaga C, Kauwe JS, Harari O, Jin SC, Cai Y, Karch CM, Benitez BA, Jeng AT, Skorupa T, Carrell D, et al. GWAS of cerebrospinal fluid tau levels identifies risk variants for Alzheimer’s disease. Neuron. 2013;78(2):256–68. doi: 10.1016/j.neuron.2013.02.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Cummings AC, Jiang L, Velez Edwards DR, McCauley JL, Laux R, McFarland LL, Fuzzell D, Knebusch C, Caywood L, Reinhart-Mercer L, et al. Genome-wide association and linkage study in the Amish detects a novel candidate late-onset Alzheimer disease gene. Ann Hum Genet. 2012;76(5):342–51. doi: 10.1111/j.1469-1809.2012.00721.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Curtis D, North BV, Sham PC. A novel method of two-locus linkage analysis applied to a genome scan for late onset Alzheimer’s disease. Ann Hum Genet. 2001;65(Pt 5):473–81. doi: 10.1017/S0003480001008776. [DOI] [PubMed] [Google Scholar]
  44. Dal Forno G, Rasmusson DX, Brandt J, Carson KA, Brookmeyer R, Troncoso J, Kawas CH. Apolipoprotein E genotype and rate of decline in probable Alzheimer’s disease. Arch Neurol. 1996;53(4):345–50. doi: 10.1001/archneur.1996.00550040085017. [DOI] [PubMed] [Google Scholar]
  45. Das HK, McPherson J, Bruns GA, Karathanasis SK, Breslow JL. Isolation, characterization, and mapping to chromosome 19 of the human apolipoprotein E gene. J Biol Chem. 1985;260(10):6240–7. [PubMed] [Google Scholar]
  46. De Strooper B. Loss-of-function presenilin mutations in Alzheimer disease. Talking Point on the role of presenilin mutations in Alzheimer disease. EMBO Rep. 2007;8(2):141–6. doi: 10.1038/sj.embor.7400897. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Deary IJ, Whiteman MC, Pattie A, Starr JM, Hayward C, Wright AF, Carothers A, Whalley LJ. Cognitive change and the APOE epsilon 4 allele. Nature. 2002;418(6901):932. doi: 10.1038/418932a. [DOI] [PubMed] [Google Scholar]
  48. Deng YL, Liu LH, Wang Y, Tang HD, Ren RJ, Xu W, Ma JF, Wang LL, Zhuang JP, Wang G, et al. The prevalence of CD33 and MS4A6A variant in Chinese Han population with Alzheimer’s disease. Hum Genet. 2012;131(7):1245–9. doi: 10.1007/s00439-012-1154-6. [DOI] [PubMed] [Google Scholar]
  49. Dik MG, Deeg DJ, Bouter LM, Corder EH, Kok A, Jonker C. Stroke and apolipoprotein E epsilon4 are independent risk factors for cognitive decline: A population-based study. Stroke. 2000;31(10):2431–6. doi: 10.1161/01.str.31.10.2431. [DOI] [PubMed] [Google Scholar]
  50. Escott-Price V, Bellenguez C, Wang LS, Choi SH, Harold D, Jones L, Holmans P, Gerrish A, Vedernikov A, Richards A, et al. Gene-wide analysis detects two new susceptibility genes for Alzheimer’s disease. PLoS One. 2014;9(6):e94661. doi: 10.1371/journal.pone.0094661. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Evans DA, Funkenstein HH, Albert MS, Scherr PA, Cook NR, Chown MJ, Hebert LE, Hennekens CH, Taylor JO. Prevalence of Alzheimer’s disease in a community population of older persons. Higher than previously reported. Jama. 1989;262(18):2551–6. [PubMed] [Google Scholar]
  52. Farrer LA, Bowirrat A, Friedland RP, Waraska K, Korczyn AD, Baldwin CT. Identification of multiple loci for Alzheimer disease in a consanguineous Israeli-Arab community. Hum Mol Genet. 2003;12(4):415–22. doi: 10.1093/hmg/ddg037. [DOI] [PubMed] [Google Scholar]
  53. Farrer LA, Cupples LA, Haines JL, Hyman B, Kukull WA, Mayeux R, Myers RH, Pericak-Vance MA, Risch N, van Duijn CM. Effects of age, sex, and ethnicity on the association between apolipoprotein E genotype and Alzheimer disease. A meta-analysis. APOE and Alzheimer Disease Meta Analysis Consortium. JAMA. 1997;278(16):1349–56. [PubMed] [Google Scholar]
  54. Farrer LA, Myers RH, Connor L, Cupples LA, Growdon JH. Segregation analysis reveals evidence of a major gene for Alzheimer disease. Am J Hum Genet. 1991;48(6):1026–33. [PMC free article] [PubMed] [Google Scholar]
  55. Farrer LA, Myers RH, Cupples LA, St George-Hyslop PH, Bird TD, Rossor MN, Mullan MJ, Polinsky R, Nee L, Heston L, et al. Transmission and age-at-onset patterns in familial Alzheimer’s disease: evidence for heterogeneity. Neurology. 1990;40(3 Pt 1):395–403. doi: 10.1212/wnl.40.3_part_1.395. [DOI] [PubMed] [Google Scholar]
  56. Fernando MS, Ince PG, Function MRCC, Ageing Neuropathology Study G Vascular pathologies and cognition in a population-based cohort of elderly people. J Neurol Sci. 2004;226(1–2):13–7. doi: 10.1016/j.jns.2004.09.004. [DOI] [PubMed] [Google Scholar]
  57. Flint J, Munafo MR. The endophenotype concept in psychiatric genetics. Psychol Med. 2007;37(2):163–80. doi: 10.1017/S0033291706008750. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Frisoni GB, Govoni S, Geroldi C, Bianchetti A, Calabresi L, Franceschini G, Trabucchi M. Gene dose of the epsilon 4 allele of apolipoprotein E and disease progression in sporadic late-onset Alzheimer’s disease. Ann Neurol. 1995;37(5):596–604. doi: 10.1002/ana.410370509. [DOI] [PubMed] [Google Scholar]
  59. Goate A, Chartier-Harlin MC, Mullan M, Brown J, Crawford F, Fidani L, Giuffra L, Haynes A, Irving N, James L, et al. Segregation of a missense mutation in the amyloid precursor protein gene with familial Alzheimer’s disease. Nature. 1991;349(6311):704–6. doi: 10.1038/349704a0. [DOI] [PubMed] [Google Scholar]
  60. Golde TE, Streit WJ, Chakrabarty P. Alzheimer’s disease risk alleles in TREM2 illuminate innate immunity in Alzheimer’s disease. Alzheimers Res Ther. 2013;5(3):24. doi: 10.1186/alzrt178. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Goldgaber D, Lerman MI, McBride OW, Saffiotti U, Gajdusek DC. Characterization and chromosomal localization of a cDNA encoding brain amyloid of Alzheimer’s disease. Science. 1987;235(4791):877–80. doi: 10.1126/science.3810169. [DOI] [PubMed] [Google Scholar]
  62. Growdon JH, Locascio JJ, Corkin S, Gomez-Isla T, Hyman BT. Apolipoprotein E genotype does not influence rates of cognitive decline in Alzheimer’s disease. Neurology. 1996;47(2):444–8. doi: 10.1212/wnl.47.2.444. [DOI] [PubMed] [Google Scholar]
  63. Grupe A, Abraham R, Li Y, Rowland C, Hollingworth P, Morgan A, Jehu L, Segurado R, Stone D, Schadt E, et al. Evidence for novel susceptibility genes for late-onset Alzheimer’s disease from a genome-wide association study of putative functional variants. Hum Mol Genet. 2007;16(8):865–73. doi: 10.1093/hmg/ddm031. [DOI] [PubMed] [Google Scholar]
  64. Gu H, Wei X, Chen S, Kurz A, Muller U, Gasser T, Dodel RC, Farlow MR, Du Y. Association of clusterin gene polymorphisms with late-onset Alzheimer’s disease. Dement Geriatr Cogn Disord. 2011;32(3):198–201. doi: 10.1159/000331276. [DOI] [PubMed] [Google Scholar]
  65. Guerreiro R, Bras J, Hardy J. SnapShot: genetics of Alzheimer’s disease. Cell. 2013a;155(4):968–968 e1. doi: 10.1016/j.cell.2013.10.037. [DOI] [PubMed] [Google Scholar]
  66. Guerreiro R, Wojtas A, Bras J, Carrasquillo M, Rogaeva E, Majounie E, Cruchaga C, Sassi C, Kauwe JS, Younkin S, et al. TREM2 variants in Alzheimer’s disease. N Engl J Med. 2013b;368(2):117–27. doi: 10.1056/NEJMoa1211851. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Guerreiro RJ, Lohmann E, Kinsella E, Bras JM, Luu N, Gurunlian N, Dursun B, Bilgic B, Santana I, Hanagasi H, et al. Exome sequencing reveals an unexpected genetic cause of disease: NOTCH3 mutation in a Turkish family with Alzheimer’s disease. Neurobiol Aging. 2012;33(5):1008 e17–23. doi: 10.1016/j.neurobiolaging.2011.10.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Hahs DW, McCauley JL, Crunk AE, McFarland LL, Gaskell PC, Jiang L, Slifer SH, Vance JM, Scott WK, Welsh-Bohmer KA, et al. A genome-wide linkage analysis of dementia in the Amish. Am J Med Genet B Neuropsychiatr Genet. 2006;141B(2):160–6. doi: 10.1002/ajmg.b.30257. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Hamshere ML, Holmans PA, Avramopoulos D, Bassett SS, Blacker D, Bertram L, Wiener H, Rochberg N, Tanzi RE, Myers A, et al. Genome-wide linkage analysis of 723 affected relative pairs with late-onset Alzheimer’s disease. Hum Mol Genet. 2007;16(22):2703–12. doi: 10.1093/hmg/ddm224. [DOI] [PubMed] [Google Scholar]
  70. Han MR, Schellenberg GD, Wang LS, Alzheimer’s Disease Neuroimaging I Genome-wide association reveals genetic effects on human Abeta42 and tau protein levels in cerebrospinal fluids: a case control study. BMC Neurol. 2010;10:90. doi: 10.1186/1471-2377-10-90. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Harold D, Abraham R, Hollingworth P, Sims R, Gerrish A, Hamshere ML, Pahwa JS, Moskvina V, Dowzell K, Williams A, et al. Genome-wide association study identifies variants at CLU and PICALM associated with Alzheimer’s disease. Nat Genet. 2009;41(10):1088–93. doi: 10.1038/ng.440. [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Hebert LE, Scherr PA, Bienias JL, Bennett DA, Evans DA. Alzheimer disease in the US population: prevalence estimates using the 2000 census. Arch Neurol. 2003;60(8):1119–22. doi: 10.1001/archneur.60.8.1119. [DOI] [PubMed] [Google Scholar]
  73. Hebert LE, Weuve J, Scherr PA, Evans DA. Alzheimer disease in the United States (2010-2050) estimated using the 2010 census. Neurology. 2013;80(19):1778–83. doi: 10.1212/WNL.0b013e31828726f5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Heilmann S, Drichel D, Clarimon J, Fernandez V, Lacour A, Wagner H, Thelen M, Hernandez I, Fortea J, Alegret M, et al. PLD3 in non-familial Alzheimer’s disease. Nature. 2015;520(7545):E3–5. doi: 10.1038/nature14039. [DOI] [PubMed] [Google Scholar]
  75. Heyman A, Wilkinson WE, Hurwitz BJ, Schmechel D, Sigmon AH, Weinberg T, Helms MJ, Swift M. Alzheimer’s disease: genetic aspects and associated clinical disorders. Ann Neurol. 1983;14(5):507–15. doi: 10.1002/ana.410140503. [DOI] [PubMed] [Google Scholar]
  76. Hilbich C, Kisters-Woike B, Reed J, Masters CL, Beyreuther K. Human and rodent sequence analogs of Alzheimer’s amyloid beta A4 share similar properties and can be solubilized in buffers of pH 7.4. Eur J Biochem. 1991;201(1):61–9. doi: 10.1111/j.1432-1033.1991.tb16256.x. [DOI] [PubMed] [Google Scholar]
  77. Hiltunen M, Mannermaa A, Thompson D, Easton D, Pirskanen M, Helisalmi S, Koivisto AM, Lehtovirta M, Ryynanen M, Soininen H. Genome-wide linkage disequilibrium mapping of late-onset Alzheimer’s disease in Finland. Neurology. 2001;57(9):1663–8. doi: 10.1212/wnl.57.9.1663. [DOI] [PubMed] [Google Scholar]
  78. Hirst C, Yee IM, Sadovnick AD. Familial risks for Alzheimer disease from a population-based series. Genet Epidemiol. 1994;11(4):365–74. doi: 10.1002/gepi.1370110406. [DOI] [PubMed] [Google Scholar]
  79. Hollingworth P, Harold D, Sims R, Gerrish A, Lambert JC, Carrasquillo MM, Abraham R, Hamshere ML, Pahwa JS, Moskvina V, et al. Common variants at ABCA7, MS4A6A/MS4A4E, EPHA1, CD33 and CD2AP are associated with Alzheimer’s disease. Nat Genet. 2011;43(5):429–35. doi: 10.1038/ng.803. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Holmans P, Hamshere M, Hollingworth P, Rice F, Tunstall N, Jones S, Moore P, Wavrant DeVrieze F, Myers A, Crook R, et al. Genome screen for loci influencing age at onset and rate of decline in late onset Alzheimer’s disease. Am J Med Genet B Neuropsychiatr Genet. 2005;135B(1):24–32. doi: 10.1002/ajmg.b.30114. [DOI] [PubMed] [Google Scholar]
  81. Hooli BV, Lill CM, Mullin K, Qiao D, Lange C, Bertram L, Tanzi RE. PLD3 gene variants and Alzheimer’s disease. Nature. 2015;520(7545):E7–8. doi: 10.1038/nature14040. [DOI] [PubMed] [Google Scholar]
  82. Hyman BT, Hedley-Whyte ET, Rebeck GW, Vonsattel JP, West HL, Growdon JH. Apolipoprotein E epsilon4/4 in a neuropathologically normal very elderly individual. Arch Neurol. 1996;53(3):215. doi: 10.1001/archneur.1996.00550030017010. [DOI] [PubMed] [Google Scholar]
  83. International HapMap C. Frazer KA, Ballinger DG, Cox DR, Hinds DA, Stuve LL, Gibbs RA, Belmont JW, Boudreau A, Hardenbol P, et al. A second generation human haplotype map of over 3.1 million SNPs. Nature. 2007;449(7164):851–61. doi: 10.1038/nature06258. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Ioannidis JP. Non-replication and inconsistency in the genome-wide association setting. Hum Hered. 2007;64(4):203–13. doi: 10.1159/000103512. [DOI] [PubMed] [Google Scholar]
  85. Irimia M, Weatheritt RJ, Ellis JD, Parikshak NN, Gonatopoulos-Pournatzis T, Babor M, Quesnel-Vallieres M, Tapial J, Raj B, O’Hanlon D, et al. A highly conserved program of neuronal microexons is misregulated in autistic brains. Cell. 2014;159(7):1511–23. doi: 10.1016/j.cell.2014.11.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Jiang T, Yu JT, Zhu XC, Tan L. TREM2 in Alzheimer’s disease. Mol Neurobiol. 2013;48(1):180–5. doi: 10.1007/s12035-013-8424-8. [DOI] [PubMed] [Google Scholar]
  87. Jin C, Li W, Yuan J, Xu W, Cheng Z. Association of the CR1 polymorphism with late-onset Alzheimer’s disease in Chinese Han populations: a meta-analysis. Neurosci Lett. 2012;527(1):46–9. doi: 10.1016/j.neulet.2012.08.032. [DOI] [PubMed] [Google Scholar]
  88. Jones L, Holmans PA, Hamshere ML, Harold D, Moskvina V, Ivanov D, Pocklington A, Abraham R, Hollingworth P, Sims R, et al. Genetic evidence implicates the immune system and cholesterol metabolism in the aetiology of Alzheimer’s disease. PLoS One. 2010;5(11):e13950. doi: 10.1371/journal.pone.0013950. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Jonsson T, Atwal JK, Steinberg S, Snaedal J, Jonsson PV, Bjornsson S, Stefansson H, Sulem P, Gudbjartsson D, Maloney J, et al. A mutation in APP protects against Alzheimer’s disease and age-related cognitive decline. Nature. 2012;488(7409):96–9. doi: 10.1038/nature11283. [DOI] [PubMed] [Google Scholar]
  90. Jonsson T, Stefansson H, Steinberg S, Jonsdottir I, Jonsson PV, Snaedal J, Bjornsson S, Huttenlocher J, Levey AI, Lah JJ, et al. Variant of TREM2 associated with the risk of Alzheimer’s disease. N Engl J Med. 2013;368(2):107–16. doi: 10.1056/NEJMoa1211103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Jun G, Naj AC, Beecham GW, Wang LS, Buros J, Gallins PJ, Buxbaum JD, Ertekin-Taner N, Fallin MD, Friedland R, et al. Meta-analysis Confirms CR1, CLU, and PICALM as Alzheimer Disease Risk Loci and Reveals Interactions With APOE Genotypes. Arch Neurol. 2010 doi: 10.1001/archneurol.2010.201. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Kamboh MI, Barmada MM, Demirci FY, Minster RL, Carrasquillo MM, Pankratz VS, Younkin SG, Saykin AJ, Sweet RA, Feingold E, et al. Genome-wide association analysis of age-at-onset in Alzheimer’s disease. Mol Psychiatry. 2012a doi: 10.1038/mp.2011.135. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Kamboh MI, Demirci FY, Wang X, Minster RL, Carrasquillo MM, Pankratz VS, Younkin SG, Saykin AJ, Alzheimer’s Disease Neuroimaging I. Jun G, et al. Genome-wide association study of Alzheimer’s disease. Transl Psychiatry. 2012b;2:e117. doi: 10.1038/tp.2012.45. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Kamboh MI, Minster RL, Demirci FY, Ganguli M, Dekosky ST, Lopez OL, Barmada MM. Association of CLU and PICALM variants with Alzheimer’s disease. Neurobiol Aging. 2012c;33(3):518–21. doi: 10.1016/j.neurobiolaging.2010.04.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Katzman R. Differential diagnosis of dementing illnesses. Neurol Clin. 1986;4(2):329–40. [PubMed] [Google Scholar]
  96. Kehoe P, Wavrant-De Vrieze F, Crook R, Wu WS, Holmans P, Fenton I, Spurlock G, Norton N, Williams H, Williams N, et al. A full genome scan for late onset Alzheimer’s disease. Hum Mol Genet. 1999;8(2):237–45. doi: 10.1093/hmg/8.2.237. [DOI] [PubMed] [Google Scholar]
  97. Kim S, Swaminathan S, Shen L, Risacher SL, Nho K, Foroud T, Shaw LM, Trojanowski JQ, Potkin SG, Huentelman MJ, et al. Genome-wide association study of CSF biomarkers Abeta1-42, t-tau, and p-tau181p in the ADNI cohort. Neurology. 2011;76(1):69–79. doi: 10.1212/WNL.0b013e318204a397. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Knopman D. Pharmacotherapy for Alzheimer’s disease: 2002. Clin Neuropharmacol. 2003;26(2):93–101. doi: 10.1097/00002826-200303000-00009. [DOI] [PubMed] [Google Scholar]
  99. Komatsu M, Shibata N, Kuerban B, Ohnuma T, Baba H, Arai H. Genetic association between clusterin polymorphisms and Alzheimer’s disease in a Japanese population. Psychogeriatrics. 2011;11(1):14–8. doi: 10.1111/j.1479-8301.2010.00346.x. [DOI] [PubMed] [Google Scholar]
  100. Kunkle BW, Jaworski J, Barral S, Vardarajan B, Beecham GW, Martin ER, Cantwell LS, Partch A, Bird TD, Raskind WH, et al. Genome-wide linkage analyses of non-Hispanic white families identify novel loci for familial late-onset Alzheimer’s disease. Alzheimers Dement. 2016;12(1):2–10. doi: 10.1016/j.jalz.2015.05.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Lambert JC, Grenier-Boley B, Bellenguez C, Pasquier F, Campion D, Dartigues JF, Berr C, Tzourio C, Amouyel P. PLD3 and sporadic Alzheimer’s disease risk. Nature. 2015;520(7545):E1. doi: 10.1038/nature14036. [DOI] [PubMed] [Google Scholar]
  102. Lambert JC, Grenier-Boley B, Harold D, Zelenika D, Chouraki V, Kamatani Y, Sleegers K, Ikram MA, Hiltunen M, Reitz C, et al. Genome-wide haplotype association study identifies the FRMD4A gene as a risk locus for Alzheimer’s disease. Mol Psychiatry. 2013a;18(4):461–70. doi: 10.1038/mp.2012.14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Lambert JC, Heath S, Even G, Campion D, Sleegers K, Hiltunen M, Combarros O, Zelenika D, Bullido MJ, Tavernier B, et al. Genome-wide association study identifies variants at CLU and CR1 associated with Alzheimer’s disease. Nat Genet. 2009;41(10):1094–9. doi: 10.1038/ng.439. [DOI] [PubMed] [Google Scholar]
  104. Lambert JC, Ibrahim-Verbaas CA, Harold D, Naj AC, Sims R, Bellenguez C, Jun G, Destefano AL, Bis JC, Beecham GW, et al. Meta-analysis of 74,046 individuals identifies 11 new susceptibility loci for Alzheimer’s disease. Nat Genet. 2013b;45(12):1452–8. doi: 10.1038/ng.2802. [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Lambert JC, Zelenika D, Hiltunen M, Chouraki V, Combarros O, Bullido MJ, Tognoni G, Fievet N, Boland A, Arosio B, et al. Evidence of the association of BIN1 and PICALM with the AD risk in contrasting European populations. Neurobiol Aging. 2011;32(4):756 e11–5. doi: 10.1016/j.neurobiolaging.2010.11.022. [DOI] [PubMed] [Google Scholar]
  106. Lautenschlager N, Kurz A, Muller U. Inheritable causes and risk factors of Alzheimer’s disease. Nervenarzt. 1999;70(3):195–205. doi: 10.1007/s001150050423. [DOI] [PubMed] [Google Scholar]
  107. Lee HB, Lyketsos CG. Depression in Alzheimer’s disease: heterogeneity and related issues. Biol Psychiatry. 2003;54(3):353–62. doi: 10.1016/s0006-3223(03)00543-2. [DOI] [PubMed] [Google Scholar]
  108. Lee JH, Barral S, Cheng R, Chacon I, Santana V, Williamson J, Lantigua R, Medrano M, Jimenez-Velazquez IZ, Stern Y, et al. Age-at-onset linkage analysis in Caribbean Hispanics with familial late-onset Alzheimer’s disease. Neurogenetics. 2008;9(1):51–60. doi: 10.1007/s10048-007-0103-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Lee JH, Cheng R, Santana V, Williamson J, Lantigua R, Medrano M, Arriaga A, Stern Y, Tycko B, Rogaeva E, et al. Expanded genomewide scan implicates a novel locus at 3q28 among Caribbean hispanics with familial Alzheimer disease. Arch Neurol. 2006;63(11):1591–8. doi: 10.1001/archneur.63.11.1591. [DOI] [PubMed] [Google Scholar]
  110. Lehtovirta M, Soininen H, Helisalmi S, Mannermaa A, Helkala EL, Hartikainen P, Hanninen T, Ryynanen M, Riekkinen PJ. Clinical and neuropsychological characteristics in familial and sporadic Alzheimer’s disease: relation to apolipoprotein E polymorphism. Neurology. 1996;46(2):413–9. doi: 10.1212/wnl.46.2.413. [DOI] [PubMed] [Google Scholar]
  111. Lemere CA, Blusztajn JK, Yamaguchi H, Wisniewski T, Saido TC, Selkoe DJ. Sequence of deposition of heterogeneous amyloid beta-peptides and APO E in Down syndrome: implications for initial events in amyloid plaque formation. Neurobiol Dis. 1996;3(1):16–32. doi: 10.1006/nbdi.1996.0003. [DOI] [PubMed] [Google Scholar]
  112. Levy-Lahad E, Wasco W, Poorkaj P, Romano DM, Oshima J, Pettingell WH, Yu CE, Jondro PD, Schmidt SD, Wang K, et al. Candidate gene for the chromosome 1 familial Alzheimer’s disease locus. Science. 1995;269(5226):973–7. doi: 10.1126/science.7638622. [DOI] [PubMed] [Google Scholar]
  113. Li YJ, Scott WK, Hedges DJ, Zhang F, Gaskell PC, Nance MA, Watts RL, Hubble JP, Koller WC, Pahwa R, et al. Age at onset in two common neurodegenerative diseases is genetically controlled. Am J Hum Genet. 2002;70(4):985–93. doi: 10.1086/339815. [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. Liu F, Arias-Vasquez A, Sleegers K, Aulchenko YS, Kayser M, Sanchez-Juan P, Feng BJ, Bertoli-Avella AM, van Swieten J, Axenovich TI, et al. A genomewide screen for late-onset Alzheimer disease in a genetically isolated Dutch population. Am J Hum Genet. 2007;81(1):17–31. doi: 10.1086/518720. [DOI] [PMC free article] [PubMed] [Google Scholar]
  115. Liu G, Wang H, Liu J, Li J, Li H, Ma G, Jiang Y, Chen Z, Zhao B, Li K. The CLU gene rs11136000 variant is significantly associated with Alzheimer’s disease in Caucasian and Asian populations. Neuromolecular Med. 2014;16(1):52–60. doi: 10.1007/s12017-013-8250-1. [DOI] [PubMed] [Google Scholar]
  116. Liu G, Zhang S, Cai Z, Li Y, Cui L, Ma G, Jiang Y, Zhang L, Feng R, Liao M, et al. BIN1 gene rs744373 polymorphism contributes to Alzheimer’s disease in East Asian population. Neurosci Lett. 2013;544:47–51. doi: 10.1016/j.neulet.2013.02.075. [DOI] [PubMed] [Google Scholar]
  117. Mahley RW. Apolipoprotein E: cholesterol transport protein with expanding role in cell biology. Science. 1988;240(4852):622–30. doi: 10.1126/science.3283935. [DOI] [PubMed] [Google Scholar]
  118. Majounie E, Abramzon Y, Renton AE, Perry R, Bassett SS, Pletnikova O, Troncoso JC, Hardy J, Singleton AB, Traynor BJ. Repeat expansion in C9ORF72 in Alzheimer’s disease. N Engl J Med. 2012;366(3):283–4. doi: 10.1056/NEJMc1113592. [DOI] [PMC free article] [PubMed] [Google Scholar]
  119. Mann DM, Iwatsubo T, Cairns NJ, Lantos PL, Nochlin D, Sumi SM, Bird TD, Poorkaj P, Hardy J, Hutton M, et al. Amyloid beta protein (Abeta) deposition in chromosome 14-linked Alzheimer’s disease: predominance of Abeta42(43) Ann Neurol. 1996;40(2):149–56. doi: 10.1002/ana.410400205. [DOI] [PubMed] [Google Scholar]
  120. Martins RN, Clarnette R, Fisher C, Broe GA, Brooks WS, Montgomery P, Gandy SE. ApoE genotypes in Australia: roles in early and late onset Alzheimer’s disease and Down’s syndrome. Neuroreport. 1995;6(11):1513–6. [PubMed] [Google Scholar]
  121. Martiskainen H, Viswanathan J, Nykanen NP, Kurki M, Helisalmi S, Natunen T, Sarajarvi T, Kurkinen KM, Pursiheimo JP, Rauramaa T, et al. Transcriptomics and mechanistic elucidation of Alzheimer’s disease risk genes in the brain and in vitro models. Neurobiol Aging. 2015;36(2):1221 e15–28. doi: 10.1016/j.neurobiolaging.2014.09.003. [DOI] [PubMed] [Google Scholar]
  122. Mayeux R. Dissecting the relative influences of genes and the environment in Alzheimer’s disease. Ann Neurol. 2004;55(2):156–8. doi: 10.1002/ana.20037. [DOI] [PubMed] [Google Scholar]
  123. Mayeux R, Lee JH, Romas SN, Mayo D, Santana V, Williamson J, Ciappa A, Rondon HZ, Estevez P, Lantigua R, et al. Chromosome-12 mapping of late-onset Alzheimer disease among Caribbean Hispanics. Am J Hum Genet. 2002;70(1):237–43. doi: 10.1086/324773. [DOI] [PMC free article] [PubMed] [Google Scholar]
  124. 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 Department of Health and Human Services Task Force on Alzheimer’s Disease. Neurology. 1984;34(7):939–44. doi: 10.1212/wnl.34.7.939. [DOI] [PubMed] [Google Scholar]
  125. Melville SA, Buros J, Parrado AR, Vardarajan B, Logue MW, Shen L, Risacher SL, Kim S, Jun G, DeCarli C, et al. Multiple loci influencing hippocampal degeneration identified by genome scan. Ann Neurol. 2012;72(1):65–75. doi: 10.1002/ana.23644. [DOI] [PMC free article] [PubMed] [Google Scholar]
  126. Missler M, Sudhof TC. Neurexophilins form a conserved family of neuropeptide-like glycoproteins. J Neurosci. 1998;18(10):3630–8. doi: 10.1523/JNEUROSCI.18-10-03630.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Miyashita A, Koike A, Jun G, Wang LS, Takahashi S, Matsubara E, Kawarabayashi T, Shoji M, Tomita N, Arai H, et al. SORL1 is genetically associated with late-onset Alzheimer’s disease in Japanese, Koreans and Caucasians. PLoS One. 2013;8(4):e58618. doi: 10.1371/journal.pone.0058618. [DOI] [PMC free article] [PubMed] [Google Scholar]
  128. Moskvina V, O’Donovan MC. Detailed analysis of the relative power of direct and indirect association studies and the implications for their interpretation. Hum Hered. 2007;64(1):63–73. doi: 10.1159/000101424. [DOI] [PubMed] [Google Scholar]
  129. Myers A, Wavrant De-Vrieze F, Holmans P, Hamshere M, Crook R, Compton D, Marshall H, Meyer D, Shears S, Booth J, et al. Full genome screen for Alzheimer disease: stage II analysis. Am J Med Genet. 2002;114(2):235–44. doi: 10.1002/ajmg.10183. [DOI] [PubMed] [Google Scholar]
  130. Nagy Z, Esiri MM, Jobst KA, Morris JH, King EM, McDonald B, Litchfield S, Smith A, Barnetson L, Smith AD. Relative roles of plaques and tangles in the dementia of Alzheimer’s disease: correlations using three sets of neuropathological criteria. Dementia. 1995;6(1):21–31. doi: 10.1159/000106918. [DOI] [PubMed] [Google Scholar]
  131. Naj AC, Beecham GW, Martin ER, Gallins PJ, Powell EH, Konidari I, Whitehead PL, Cai G, Haroutunian V, Scott WK, et al. Dementia Revealed: Novel Chromosome 6 Locus for Late-Onset Alzheimer Disease Provides Genetic Evidence for Folate-Pathway Abnormalities. PLoS Genet. 2010;6(9) doi: 10.1371/journal.pgen.1001130. [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. Naj AC, Jun G, Beecham GW, Wang LS, Vardarajan BN, Buros J, Gallins PJ, Buxbaum JD, Jarvik GP, Crane PK, et al. Common variants at MS4A4/MS4A6E, CD2AP, CD33 and EPHA1 are associated with late-onset Alzheimer’s disease. Nat Genet. 2011;43(5):436–41. doi: 10.1038/ng.801. [DOI] [PMC free article] [PubMed] [Google Scholar]
  133. Naj AC, Jun G, Reitz C, Kunkle BW, Perry W, Park YS, Beecham GW, Rajbhandary RA, Hamilton-Nelson KL, Wang LS, et al. Effects of multiple genetic loci on age at onset in late-onset Alzheimer disease: a genome-wide association study. JAMA Neurol. 2014;71(11):1394–404. doi: 10.1001/jamaneurol.2014.1491. [DOI] [PMC free article] [PubMed] [Google Scholar]
  134. Neale BM, Sham PC. The future of association studies: gene-based analysis and replication. Am J Hum Genet. 2004;75(3):353–62. doi: 10.1086/423901. [DOI] [PMC free article] [PubMed] [Google Scholar]
  135. Nee LE, Polinsky RJ, Eldridge R, Weingartner H, Smallberg S, Ebert M. A family with histologically confirmed Alzheimer’s disease. Arch Neurol. 1983;40(4):203–8. doi: 10.1001/archneur.1983.04050040033004. [DOI] [PubMed] [Google Scholar]
  136. Olaisen B, Teisberg P, Gedde-Dahl T., Jr The locus for apolipoprotein E (apoE) is linked to the complement component C3 (C3) locus on chromosome 19 in man. Hum Genet. 1982;62(3):233–6. doi: 10.1007/BF00333526. [DOI] [PubMed] [Google Scholar]
  137. Olin JT, Schneider LS, Katz IR, Meyers BS, Alexopoulos GS, Breitner JC, Bruce ML, Caine ED, Cummings JL, Devanand DP, et al. Provisional diagnostic criteria for depression of Alzheimer disease. Am J Geriatr Psychiatry. 2002;10(2):125–8. [PubMed] [Google Scholar]
  138. Olson JM, Goddard KA, Dudek DM. A second locus for very-late-onset Alzheimer disease: a genome scan reveals linkage to 20p and epistasis between 20p and the amyloid precursor protein region. Am J Hum Genet. 2002;71(1):154–61. doi: 10.1086/341034. [DOI] [PMC free article] [PubMed] [Google Scholar]
  139. Omoumi A, Fok A, Greenwood T, Sadovnick AD, Feldman HH, Hsiung GY. Evaluation of late-onset Alzheimer disease genetic susceptibility risks in a Canadian population. Neurobiol Aging. 2014;35(4):936 e5–12. doi: 10.1016/j.neurobiolaging.2013.09.025. [DOI] [PubMed] [Google Scholar]
  140. Oyama F, Shimada H, Oyama R, Ihara Y. Apolipoprotein E genotype, Alzheimer’s pathologies and related gene expression in the aged population. Brain Res Mol Brain Res. 1995;29(1):92–8. doi: 10.1016/0169-328x(94)00233-5. [DOI] [PubMed] [Google Scholar]
  141. Paloneva J, Manninen T, Christman G, Hovanes K, Mandelin J, Adolfsson R, Bianchin M, Bird T, Miranda R, Salmaggi A, et al. Mutations in two genes encoding different subunits of a receptor signaling complex result in an identical disease phenotype. Am J Hum Genet. 2002;71(3):656–62. doi: 10.1086/342259. [DOI] [PMC free article] [PubMed] [Google Scholar]
  142. Pericak-Vance MA, Bass MP, Yamaoka LH, Gaskell PC, Scott WK, Terwedow HA, Menold MM, Conneally PM, Small GW, Vance JM, et al. Complete genomic screen in late-onset familial Alzheimer disease. Evidence for a new locus on chromosome 12. JAMA. 1997;278(15):1237–41. [PubMed] [Google Scholar]
  143. Pericak-Vance MA, Bebout JL, Gaskell PC, Jr, Yamaoka LH, Hung WY, Alberts MJ, Walker AP, Bartlett RJ, Haynes CA, Welsh KA, et al. Linkage studies in familial Alzheimer disease: evidence for chromosome 19 linkage. Am J Hum Genet. 1991;48(6):1034–50. [PMC free article] [PubMed] [Google Scholar]
  144. Pericak-Vance MA, Conneally PM, Gaskell PC, Small GW, Saunders AM, Yamaoka LH, Robinson C, Ter-Minassian M, Locker PA, Pritchard M, et al. The search for additional late-onset Alzheimer disease gene: a complete genomic screen. Journal of Human Genetics. 1995;57:A200. [Google Scholar]
  145. Pericak-Vance MA, Grubber J, Bailey LR, Hedges D, West S, Santoro L, Kemmerer B, Hall JL, Saunders AM, Roses AD, et al. Identification of novel genes in late-onset Alzheimer’s disease. Exp Gerontol. 2000;35(9–10):1343–52. doi: 10.1016/s0531-5565(00)00196-0. [DOI] [PubMed] [Google Scholar]
  146. Pericak-Vance MA, Yamaoka LH, Haynes CS, Speer MC, Haines JL, Gaskell PC, Hung WY, Clark CM, Heyman AL, Trofatter JA, et al. Genetic linkage studies in Alzheimer’s disease families. Exp Neurol. 1988;102(3):271–9. doi: 10.1016/0014-4886(88)90220-8. [DOI] [PubMed] [Google Scholar]
  147. Piaceri I, Bagnoli S, Lucenteforte E, Mancuso M, Tedde A, Siciliano G, Piacentini S, Bracco L, Sorbi S, Nacmias B. Implication of a genetic variant at PICALM in Alzheimer’s disease patients and centenarians. J Alzheimers Dis. 2011;24(3):409–13. doi: 10.3233/JAD-2011-101791. [DOI] [PubMed] [Google Scholar]
  148. Porter VR, Buxton WG, Fairbanks LA, Strickland T, O’Connor SM, Rosenberg-Thompson S, Cummings JL. Frequency and characteristics of anxiety among patients with Alzheimer’s disease and related dementias. J Neuropsychiatry Clin Neurosci. 2003;15(2):180–6. doi: 10.1176/jnp.15.2.180. [DOI] [PubMed] [Google Scholar]
  149. Potkin SG, Guffanti G, Lakatos A, Turner JA, Kruggel F, Fallon JH, Saykin AJ, Orro A, Lupoli S, Salvi E, et al. Hippocampal atrophy as a quantitative trait in a genome-wide association study identifying novel susceptibility genes for Alzheimer’s disease. PLoS One. 2009;4(8):e6501. doi: 10.1371/journal.pone.0006501. [DOI] [PMC free article] [PubMed] [Google Scholar]
  150. Pottier C, Hannequin D, Coutant S, Rovelet-Lecrux A, Wallon D, Rousseau S, Legallic S, Paquet C, Bombois S, Pariente J, et al. High frequency of potentially pathogenic SORL1 mutations in autosomal dominant early-onset Alzheimer disease. Mol Psychiatry. 2012;17(9):875–9. doi: 10.1038/mp.2012.15. [DOI] [PubMed] [Google Scholar]
  151. Prasher VP, Farrer MJ, Kessling AM, Fisher EM, West RJ, Barber PC, Butler AC. Molecular mapping of Alzheimer-type dementia in Down’s syndrome. Ann Neurol. 1998;43(3):380–3. doi: 10.1002/ana.410430316. [DOI] [PubMed] [Google Scholar]
  152. Prencipe M, Casini AR, Ferretti C, Lattanzio MT, Fiorelli M, Culasso F. Prevalence of dementia in an elderly rural population: effects of age, sex, and education. J Neurol Neurosurg Psychiatry. 1996;60(6):628–33. doi: 10.1136/jnnp.60.6.628. [DOI] [PMC free article] [PubMed] [Google Scholar]
  153. Rall SC, Jr, Weisgraber KH, Mahley RW. Human apolipoprotein E. The complete amino acid sequence. J Biol Chem. 1982;257(8):4171–8. [PubMed] [Google Scholar]
  154. Ramanan VK, Kim S, Holohan K, Shen L, Nho K, Risacher SL, Foroud TM, Mukherjee S, Crane PK, Aisen PS, et al. Genome-wide pathway analysis of memory impairment in the Alzheimer’s Disease Neuroimaging Initiative (ADNI) cohort implicates gene candidates, canonical pathways, and networks. Brain Imaging Behav. 2012;6(4):634–48. doi: 10.1007/s11682-012-9196-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  155. Rao VS, van Duijn CM, Connor-Lacke L, Cupples LA, Growdon JH, Farrer LA. Multiple etiologies for Alzheimer disease are revealed by segregation analysis. Am J Hum Genet. 1994;55(5):991–1000. [PMC free article] [PubMed] [Google Scholar]
  156. Rebeck GW, Reiter JS, Strickland DK, Hyman BT. Apolipoprotein E in sporadic Alzheimer’s disease: allelic variation and receptor interactions. Neuron. 1993;11(4):575–80. doi: 10.1016/0896-6273(93)90070-8. [DOI] [PubMed] [Google Scholar]
  157. Reiman EM, Webster JA, Myers AJ, Hardy J, Dunckley T, Zismann VL, Joshipura KD, Pearson JV, Hu-Lince D, Huentelman MJ, et al. GAB2 alleles modify Alzheimer’s risk in APOE epsilon4 carriers. Neuron. 2007;54(5):713–20. doi: 10.1016/j.neuron.2007.05.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  158. Reitz C, Jun G, Naj A, Rajbhandary R, Vardarajan BN, Wang LS, Valladares O, Lin CF, Larson EB, Graff-Radford NR, et al. Variants in the ATP-binding cassette transporter (ABCA7), apolipoprotein E 4, and the risk of late-onset Alzheimer disease in African Americans. JAMA. 2013;309(14):1483–92. doi: 10.1001/jama.2013.2973. [DOI] [PMC free article] [PubMed] [Google Scholar]
  159. Rice JP, Saccone NL, Rasmussen E. Definition of the phenotype. Adv Genet. 2001;42:69–76. doi: 10.1016/s0065-2660(01)42015-3. [DOI] [PubMed] [Google Scholar]
  160. Rocca WA, Hofman A, Brayne C, Breteler MM, Clarke M, Copeland JR, Dartigues JF, Engedal K, Hagnell O, Heeren TJ, et al. Frequency and distribution of Alzheimer’s disease in Europe: a collaborative study of 1980-1990 prevalence findings. The EURODEM-Prevalence Research Group. Ann Neurol. 1991;30(3):381–90. doi: 10.1002/ana.410300310. [DOI] [PubMed] [Google Scholar]
  161. Rogaev EI, Sherrington R, Rogaeva EA, Levesque G, Ikeda M, Liang Y, Chi H, Lin C, Holman K, Tsuda T, et al. Familial Alzheimer’s disease in kindreds with missense mutations in a gene on chromosome 1 related to the Alzheimer’s disease type 3 gene. Nature. 1995;376(6543):775–8. doi: 10.1038/376775a0. [DOI] [PubMed] [Google Scholar]
  162. Rogaeva E, Meng Y, Lee JH, Gu Y, Kawarai T, Zou F, Katayama T, Baldwin CT, Cheng R, Hasegawa H, et al. The neuronal sortilin-related receptor SORL1 is genetically associated with Alzheimer disease. Nat Genet. 2007;39(2):168–77. doi: 10.1038/ng1943. [DOI] [PMC free article] [PubMed] [Google Scholar]
  163. Ropacki SA, Jeste DV. Epidemiology of and risk factors for psychosis of Alzheimer’s disease: a review of 55 studies published from 1990 to 2003. Am J Psychiatry. 2005;162(11):2022–30. doi: 10.1176/appi.ajp.162.11.2022. [DOI] [PubMed] [Google Scholar]
  164. Roses AD, Devlin B, Conneally PM, Small GW, Saunders AM, Pritchard M, Locke PA, Haines JL, Pericak-Vance MA, Risch N. Measuring the genetic contribution of APOE in late-onset Alzheimer disease (AD) American Journal of Human Genetics. 1995;57:A202. [Google Scholar]
  165. Ruiz A, Heilmann S, Becker T, Hernandez I, Wagner H, Thelen M, Mauleon A, Rosende-Roca M, Bellenguez C, Bis JC, et al. Follow-up of loci from the International Genomics of Alzheimer’s Disease Project identifies TRIP4 as a novel susceptibility gene. Transl Psychiatry. 2014;4:e358. doi: 10.1038/tp.2014.2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  166. Sadovnick AD, Irwin ME, Baird PA, Beattie BL. Genetic studies on an Alzheimer clinic population. Genet Epidemiol. 1989;6(5):633–43. doi: 10.1002/gepi.1370060508. [DOI] [PubMed] [Google Scholar]
  167. Saunders AM, Strittmatter WJ, Schmechel D, George-Hyslop PH, Pericak-Vance MA, Joo SH, Rosi BL, Gusella JF, Crapper-MacLachlan DR, Alberts MJ, et al. Association of apolipoprotein E allele epsilon 4 with late-onset familial and sporadic Alzheimer’s disease. Neurology. 1993;43(8):1467–72. doi: 10.1212/wnl.43.8.1467. [DOI] [PubMed] [Google Scholar]
  168. Schellenberg GD, Bird TD, Wijsman EM, Orr HT, Anderson L, Nemens E, White JA, Bonnycastle L, Weber JL, Alonso ME, et al. Genetic linkage evidence for a familial Alzheimer’s disease locus on chromosome 14. Science. 1992;258(5082):668–71. doi: 10.1126/science.1411576. [DOI] [PubMed] [Google Scholar]
  169. Schellenberg GD, Deeb SS, Boehnke M, Bryant EM, Martin GM, Lampe TH, Bird TD. Association of an apolipoprotein CII allele with familial dementia of the Alzheimer type. J Neurogenet. 1987;4(2–3):97–108. [PubMed] [Google Scholar]
  170. Schleinitz N, Chiche L, Guia S, Bouvier G, Vernier J, Morice A, Houssaint E, Harle JR, Kaplanski G, Montero-Julian FA, et al. Pattern of DAP12 expression in leukocytes from both healthy and systemic lupus erythematosus patients. PLoS One. 2009;4(7):e6264. doi: 10.1371/journal.pone.0006264. [DOI] [PMC free article] [PubMed] [Google Scholar]
  171. Schmechel DE, Saunders AM, Strittmatter WJ, Crain BJ, Hulette CM, Joo SH, Pericak-Vance MA, Goldgaber D, Roses AD. Increased amyloid beta-peptide deposition in cerebral cortex as a consequence of apolipoprotein E genotype in late-onset Alzheimer disease. Proc Natl Acad Sci U S A. 1993;90(20):9649–53. doi: 10.1073/pnas.90.20.9649. [DOI] [PMC free article] [PubMed] [Google Scholar]
  172. Schneider JA, Arvanitakis Z, Bang W, Bennett DA. Mixed brain pathologies account for most dementia cases in community-dwelling older persons. Neurology. 2007;69(24):2197–204. doi: 10.1212/01.wnl.0000271090.28148.24. [DOI] [PubMed] [Google Scholar]
  173. Scott WK, Hauser ER, Schmechel DE, Welsh-Bohmer KA, Small GW, Roses AD, Saunders AM, Gilbert JR, Vance JM, Haines JL, et al. Ordered-subsets linkage analysis detects novel Alzheimer disease loci on chromosomes 2q34 and 15q22. Am J Hum Genet. 2003;73(5):1041–51. doi: 10.1086/379083. [DOI] [PMC free article] [PubMed] [Google Scholar]
  174. Selkoe DJ. Presenilin, Notch, and the genesis and treatment of Alzheimer’s disease. Proc Natl Acad Sci U S A. 2001;98(20):11039–41. doi: 10.1073/pnas.211352598. [DOI] [PMC free article] [PubMed] [Google Scholar]
  175. Seshadri S, Fitzpatrick AL, Ikram MA, DeStefano AL, Gudnason V, Boada M, Bis JC, Smith AV, Carassquillo MM, Lambert JC, et al. Genome-wide analysis of genetic loci associated with Alzheimer disease. Jama. 2010;303(18):1832–40. doi: 10.1001/jama.2010.574. [DOI] [PMC free article] [PubMed] [Google Scholar]
  176. Shen L, Kim S, Risacher SL, Nho K, Swaminathan S, West JD, Foroud T, Pankratz N, Moore JH, Sloan CD, et al. Whole genome association study of brain-wide imaging phenotypes for identifying quantitative trait loci in MCI and AD: A study of the ADNI cohort. Neuroimage. 2010;53(3):1051–63. doi: 10.1016/j.neuroimage.2010.01.042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  177. Sherrington R, Rogaev EI, Liang Y, Rogaeva EA, Levesque G, Ikeda M, Chi H, Lin C, Li G, Holman K, et al. Cloning of a gene bearing missense mutations in early-onset familial Alzheimer’s disease. Nature. 1995;375(6534):754–60. doi: 10.1038/375754a0. [DOI] [PubMed] [Google Scholar]
  178. Sillen A, Andrade J, Lilius L, Forsell C, Axelman K, Odeberg J, Winblad B, Graff C. Expanded high-resolution genetic study of 109 Swedish families with Alzheimer’s disease. Eur J Hum Genet. 2008;16(2):202–8. doi: 10.1038/sj.ejhg.5201946. [DOI] [PubMed] [Google Scholar]
  179. Sjogren T, Sjogren H, Lindgren AG. Morbus Alzheimer and morbus Pick; a genetic, clinical and patho-anatomical study. Acta Psychiatr Neurol Scand Suppl. 1952;82:1–152. [PubMed] [Google Scholar]
  180. St George-Hyslop P, Haines J, Rogaev E, Mortilla M, Vaula G, Pericak-Vance M, Foncin JF, Montesi M, Bruni A, Sorbi S, et al. Genetic evidence for a novel familial Alzheimer’s disease locus on chromosome 14. Nat Genet. 1992;2(4):330–4. doi: 10.1038/ng1292-330. [DOI] [PubMed] [Google Scholar]
  181. St George-Hyslop PH, Myers RH, Haines JL, Farrer LA, Tanzi RE, Abe K, James MF, Conneally PM, Polinsky RJ, Gusella JF. Familial Alzheimer’s disease: progress and problems. Neurobiol Aging. 1989;10(5):417–25. doi: 10.1016/0197-4580(89)90082-1. [DOI] [PubMed] [Google Scholar]
  182. Stein JL, Hua X, Lee S, Ho AJ, Leow AD, Toga AW, Saykin AJ, Shen L, Foroud T, Pankratz N, et al. Voxelwise genome-wide association study (vGWAS) Neuroimage. 2010;53(3):1160–74. doi: 10.1016/j.neuroimage.2010.02.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
  183. Stern Y, Brandt J, Albert M, Jacobs DM, Liu X, Bell K, Marder K, Sano M, Albert S, Del-Castillo Castenada C, et al. The absence of an apolipoprotein epsilon4 allele is associated with a more aggressive form of Alzheimer’s disease. Ann Neurol. 1997;41(5):615–20. doi: 10.1002/ana.410410510. [DOI] [PubMed] [Google Scholar]
  184. Strittmatter WJ, Weisgraber KH, Huang DY, Dong LM, Salvesen GS, Pericak-Vance M, Schmechel D, Saunders AM, Goldgaber D, Roses AD. Binding of human apolipoprotein E to synthetic amyloid beta peptide: isoform-specific effects and implications for late-onset Alzheimer disease. Proc Natl Acad Sci U S A. 1993;90(17):8098–102. doi: 10.1073/pnas.90.17.8098. [DOI] [PMC free article] [PubMed] [Google Scholar]
  185. Suzuki N, Cheung TT, Cai XD, Odaka A, Otvos L, Jr, Eckman C, Golde TE, Younkin SG. An increased percentage of long amyloid beta protein secreted by familial amyloid beta protein precursor (beta APP717) mutants. Science. 1994;264(5163):1336–40. doi: 10.1126/science.8191290. [DOI] [PubMed] [Google Scholar]
  186. Tan L, Yu JT, Zhang W, Wu ZC, Zhang Q, Liu QY, Wang W, Wang HF, Ma XY, Cui WZ. Association of GWAS-linked loci with late-onset Alzheimer’s disease in a northern Han Chinese population. Alzheimers Dement. 2013;9(5):546–53. doi: 10.1016/j.jalz.2012.08.007. [DOI] [PubMed] [Google Scholar]
  187. Tanzi RE, Bird ED, Latt SA, Neve RL. The amyloid beta protein gene is not duplicated in brains from patients with Alzheimer’s disease. Science. 1987;238(4827):666–9. doi: 10.1126/science.2890207. [DOI] [PubMed] [Google Scholar]
  188. Teri L, Ferretti LE, Gibbons LE, Logsdon RG, McCurry SM, Kukull WA, McCormick WC, Bowen JD, Larson EB. Anxiety of Alzheimer’s disease: prevalence, and comorbidity. J Gerontol A Biol Sci Med Sci. 1999;54(7):M348–52. doi: 10.1093/gerona/54.7.m348. [DOI] [PubMed] [Google Scholar]
  189. van Bon BW, Koolen DA, Brueton L, McMullan D, Lichtenbelt KD, Ades LC, Peters G, Gibson K, Moloney S, Novara F, et al. The 2q23.1 microdeletion syndrome: clinical and behavioural phenotype. Eur J Hum Genet. 2010;18(2):163–70. doi: 10.1038/ejhg.2009.152. [DOI] [PMC free article] [PubMed] [Google Scholar]
  190. Van Broeckhoven C, Backhovens H, Cruts M, Martin JJ, Crook R, Houlden H, Hardy J. APOE genotype does not modulate age of onset in families with chromosome 14 encoded Alzheimer’s disease. Neurosci Lett. 1994;169(1–2):179–80. doi: 10.1016/0304-3940(94)90385-9. [DOI] [PubMed] [Google Scholar]
  191. van der Lee SJ, Holstege H, Wong TH, Jakobsdottir J, Bis JC, Chouraki V, van Rooij JG, Grove ML, Smith AV, Amin N, et al. PLD3 variants in population studies. Nature. 2015;520(7545):E2–3. doi: 10.1038/nature14038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  192. van Duijn CM, Farrer LA, Cupples LA, Hofman A. Genetic transmission of Alzheimer’s disease among families in a Dutch population based study. J Med Genet. 1993;30(8):640–6. doi: 10.1136/jmg.30.8.640. [DOI] [PMC free article] [PubMed] [Google Scholar]
  193. Wang K, Li M, Bucan M. Pathway-based approaches for analysis of genomewide association studies. Am J Hum Genet. 2007;81(6):1278–83. doi: 10.1086/522374. [DOI] [PMC free article] [PubMed] [Google Scholar]
  194. Weisgraber KH, Innerarity TL, Mahley RW. Abnormal lipoprotein receptor-binding activity of the human E apoprotein due to cysteine-arginine interchange at a single site. J Biol Chem. 1982;257(5):2518–21. [PubMed] [Google Scholar]
  195. Welsh-Bohmer KA, Gearing M, Saunders AM, Roses AD, Mirra S. Apolipoprotein E genotypes in a neuropathological series from the Consortium to Establish a Registry for Alzheimer’s Disease. Ann Neurol. 1997;42(3):319–25. doi: 10.1002/ana.410420308. [DOI] [PubMed] [Google Scholar]
  196. Yu JT, Ma XY, Wang YL, Sun L, Tan L, Hu N, Tan L. Genetic variation in clusterin gene and Alzheimer’s disease risk in Han Chinese. Neurobiol Aging. 2013;34(7):1921 e17–23. doi: 10.1016/j.neurobiolaging.2013.01.010. [DOI] [PubMed] [Google Scholar]
  197. Zhang B, Gaiteri C, Bodea LG, Wang Z, McElwee J, Podtelezhnikov AA, Zhang C, Xie T, Tran L, Dobrin R, et al. Integrated systems approach identifies genetic nodes and networks in late-onset Alzheimer’s disease. Cell. 2013;153(3):707–20. doi: 10.1016/j.cell.2013.03.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  198. Zubenko GS, Stiffler S, Stabler S, Kopp U, Hughes HB, Cohen BM, Moossy J. Association of the apolipoprotein E epsilon 4 allele with clinical subtypes of autopsy-confirmed Alzheimer’s disease. Am J Med Genet. 1994;54(3):199–205. doi: 10.1002/ajmg.1320540306. [DOI] [PubMed] [Google Scholar]

RESOURCES