Skip to main content
American Journal of Neurodegenerative Disease logoLink to American Journal of Neurodegenerative Disease
. 2013 Sep 18;2(3):145–175.

Pathways to neurodegeneration: mechanistic insights from GWAS in Alzheimer’s disease, Parkinson’s disease, and related disorders

Vijay K Ramanan 1,2,3, Andrew J Saykin 1,2,4,5
PMCID: PMC3783830  PMID: 24093081

Abstract

The discovery of causative genetic mutations in affected family members has historically dominated our understanding of neurodegenerative diseases such as Alzheimer’s disease (AD), Parkinson’s disease (PD), frontotemporal dementia (FTD), and amyotrophic lateral sclerosis (ALS). Nevertheless, most cases of neurodegenerative disease are not explained by Mendelian inheritance of known genetic variants, but instead are thought to have a complex etiology with numerous genetic and environmental factors contributing to susceptibility. Although unbiased genome-wide association studies (GWAS) have identified novel associations to neurodegenerative diseases, most of these hits explain only modest fractions of disease heritability. In addition, despite the substantial overlap of clinical and pathologic features among major neurodegenerative diseases, surprisingly few GWAS-implicated variants appear to exhibit cross-disease association. These realities suggest limitations of the focus on individual genetic variants and create challenges for the development of diagnostic and therapeutic strategies, which traditionally target an isolated molecule or mechanistic step. Recently, GWAS of complex diseases and traits have focused less on individual susceptibility variants and instead have emphasized the biological pathways and networks revealed by genetic associations. This new paradigm draws on the hypothesis that fundamental disease processes may be influenced on a personalized basis by a combination of variants – some common and others rare, some protective and others deleterious – in key genes and pathways. Here, we review and synthesize the major pathways implicated in neurodegeneration, focusing on GWAS from the most prevalent neurodegenerative disorders, AD and PD. Using literature mining, we also discover a novel regulatory network that is enriched with AD- and PD-associated genes and centered on the SP1 and AP-1 (Jun/Fos) transcription factors. Overall, this pathway- and network-driven model highlights several potential shared mechanisms in AD and PD that will inform future studies of these and other neurodegenerative disorders. These insights also suggest that biomarker and treatment strategies may require simultaneous targeting of multiple components, including some specific to disease stage, in order to assess and modulate neurodegeneration. Pathways and networks will provide ideal vehicles for integrating relevant findings from GWAS and other modalities to enhance clinical translation.

Keywords: Neurodegeneration, Alzheimer’s disease (AD), Parkinson’s disease (PD), genome-wide association study (GWAS), single nucleotide polymorphism (SNP), pathway, network, biomarker, omics, complex disease

Introduction

Several common themes have driven prevailing notions about neurodegenerative diseases and their underlying etiology. Pathologically, a frequent characteristic of these diseases is the accumulation and aggregation of abnormal or misfolded proteins, as with amyloid-β (Aβ) in Alzheimer’s disease (AD) [1,2], α-synuclein in Parkinson’s disease (PD) [3], huntingtin protein in Huntington’s disease (HD) [4], and transactive response DNA-binding protein 43 (TDP-43) in frontotemporal dementia (FTD) and amyotrophic lateral sclerosis (ALS) [5]. The discovery of genetic mutations causing rare, early onset, familial forms of these diseases, as with the APP (amyloid precursor protein) gene in AD [6] and the SNCA (α-synuclein) gene in PD [7], further focused attention on mechanisms directly connected to disease pathology. However, most cases of AD, PD, and other neurodegenerative diseases cannot be explained by simple Mendelian inheritance of genetic mutations in isolated disease-specific pathways. These late onset, sporadic forms of disease are thought instead to have a complex etiology, with susceptibility influenced by lifestyle and environmental factors in addition to as-yet-uncharacterized variants in numerous genes [8-12].

The development of methods for unbiased investigation of the genome initially promised to address this knowledge gap. Although analyses of neurodegenerative diseases represent a substantial fraction of the more than 1500 published genome-wide association studies (GWAS) [13], several limitations of this approach have emerged. Most GWAS-implicated common single nucleotide polymorphisms (SNPs) display modest individual effects on disease risk and together leave substantial heritability unexplained [11]. For example, although up to 60-80% of AD risk is estimated to derive from genetic factors [14], known genes including the uniquely large effect of APOE (apolipoprotein E) account for just half of this genetic variance [15]. In addition, while major psychiatric disorders have displayed genetic overlap through GWAS [16], similarly robust cross-disorder SNP associations have not been reported for neurodegenerative diseases, a surprising finding given their vast overlap of clinical and pathological features. As a result, there has been significant interest in the development of alternative perspectives and analytical strategies to better understand the genetic architecture underlying neurodegeneration [17,18].

Recently, biological pathways and networks have become focal points for harnessing GWAS data [19,20]. Numerous studies have demonstrated that genes functioning in the same pathway can collectively influence susceptibility to neurodegenerative diseases and traits, even when constituent SNPs do not individually exhibit significant association [21-28]. Pathways occupied by top GWAS “hits” can also highlight additional genes with more modest effects on disease risk but which may provide better targets for biomarker and drug development [29,30]. Further, GWAS-implicated pathways and networks provide mechanistic hypotheses which can guide confirmatory testing in independent human study datasets, cell lines, and animal models. The ability to prioritize pathways of interest may be particularly important for approaches with high computational demand. These include whole genome sequencing (WGS) studies, which offer enhanced power to detect rare SNPs and copy number and other structural variants [31], studies of disease endophenotypes such as brain imaging [32,33] or cerebrospinal fluid (CSF) biomarkers [34,35], and studies of molecular interactions and epistasis [36-38], among other approaches.

We propose that pathways and networks provide an ideal framework for integrating known neurodegenerative mechanisms and nominating new targets. Here, we review the major pathways influencing neurodegeneration, focusing on shared processes implicated by GWAS of the most prevalent neurodegenerative disorders, AD and PD [39]. As part of a conceptual model (Figure 1), we discuss these pathways within broader, biologically driven groups representing intracellular mechanisms, influences from the local tissue environment and systemic circulation, and broader factors related to neurodevelopment and aging. We also perform network analysis of top AD- and PD-associated genes to discover additional novel functional relationships among multiple candidate genes and pathways.

Figure 1.

Figure 1

Conceptual model of candidate pathways contributing to neurodegeneration. Candidate pathways influencing the balance of neuronal survival and degeneration are displayed within broader functional groups based on their major site or mode of action (intracellular mechanisms, local tissue environment influences, systemic influences, and mechanisms related to neurodevelopment and aging). The pathways and overarching functional groups in this model are highly related and can have overlapping or interacting components which can collectively modulate neurodegenerative processes.

Intracellular mechanisms

Apoptosis

Although definitions vary, apoptosis is generally understood as a programmed cell death process involving caspase activation, maintenance of organelle integrity, and a lack of cell swelling [40]. Aberrant regulation of apoptosis is one proposed explanation for the striking loss of hippocampal and cortical neurons in AD and midbrain dopaminergic neurons in PD [41]. In cultured neurons, Aβ deposition is a direct inducer of apoptosis [42], and early onset AD-associated mutations in the Aβ processing genes APP, PSEN1 (presenilin-1), and PSEN2 (presenilin-2) can promote apoptosis [43-45]. The largest known genetic risk factor for lateonset AD, the APOE ε4 allele [46], may also be related to apoptosis through interactions with Aβ [47]. Interestingly, a recent protein interaction network analysis identified PDCD4 (programmed cell death 4), which is up-regulated in AD brains and whose expressed protein interacts with ApoE and presenilin-2, as a potential regulator of neuronal death in AD that may bridge genetic risk factors for early- and late-onset disease [48].

Several major AD GWAS-implicated genes [49-51] also have putative roles in apoptosis. CLU (clusterin) is proposed to interact with BCL-2 protein family members to regulate apoptosis [52,53], and neuroimaging studies suggest CLU-associated brain atrophy may be particularly evident in early stages of disease [54]. Another BCL-2 interacting gene, HRK (harakiri) [55], was identified in a large GWAS meta-analysis of magnetic resonance imaging (MRI) hippocampal volume, a key AD endophenotype [56]. Sequence homology and functional studies also indicate that ABCA7 (ATP-binding cassette transporter A7) is required for efficient clearance of apoptotic cells [57]. These diverse roles suggest that AD-associated genetic variation may have pleiotropic influences on apoptotic mechanisms.

In human PD brains, molecular markers of apoptosis are abundant in the substantia nigra [58], which contains mostly dopaminergic neurons and is the primary site of atrophy and pathology in the disease [39]. The hallmark pathological feature of PD is the presence of intracellular inclusions known as Lewy bodies, which are mainly composed of insoluble aggregates of α-synuclein [3]. SNCA is associated with both early- and late-onset PD [7,59] and accumulation of α-synuclein in cultured dopaminergic neurons results in apoptosis [60]. Other PD-related genes with potential roles in apoptosis include LRRK2 (leucine-rich repeat kinase 2) [61,62], MAPT (microtubule-associated protein tau) [63], and PARK2 (parkinson protein 2, E3 ubiquitin protein ligase) [64].

Development of anti-apoptotic and other neuroprotective drugs for AD and PD is still in early stages and may ultimately require targeting of multiple genes and sub-pathways [65,66]. Such therapies will also need to address the evolving understanding of epidemiologic and mechanistic relationships between neurodegeneration and cancer, particularly since many cancers are marked by down-regulation of apoptosis in contrast to the up-regulation seen in neurodegeneration [67,68]. Nevertheless, the heavy footprint of apoptotic functions among known AD and PD risk loci is encouraging for this direction of study.

Autophagy

Autophagy is a highly regulated mechanism for degradation of unnecessary or dysfunctional cellular components [4]. Controlled activation of autophagy may provide a strategy for clearance of long-lived, aggregated, or dysfunctional proteins which contribute to neurodegeneration [40]. In human brains, autophagy is transcriptionally down-regulated during normal aging but is up-regulated in AD, suggesting a possible attempted compensatory response to Aβ accumulation [69]. In mice, deletion of PD-related LRRK2 yields impaired autophagy and augmented accumulation of α-synuclein [62]. Variants in GBA (glucosidase-β, acid) are also associated with PD [59,70], and the accumulation of α-synuclein in mutant GBA cell lines can be reversed with administration of the autophagy inducer rapamycin [71].

An important caveat of these findings is that other potentially related outcome measures may be relevant for interpretation. For example, increased levels of apoptotic effectors such as caspase-3 have been detected after pharmacologic inhibition of autophagy in an AD mouse model [72]. Whether this represents possible cross-talk between autophagy and apoptosis to respond to cellular stress or indicates that autophagy itself is an alternate mechanism for programmed cell death remains controversial [40,73]. Genetic analyses for epistasis (gene-gene interactions) within and between these pathways may provide alternative strategies for addressing these issues. Nevertheless, the potential for complex relationships between autophagic and other pathways involved in protein stress and response suggest that in vivo modulation of autophagy as a therapy for neurodegeneration may require fine-tuning to broader genetic and environmental profiles [69,74].

Mitochondrial dysfunction

Mitochondria are primarily tasked with cellular energy production through catabolism of sugars, fats, and proteins. The underlying mechanisms for these processes are well-known to yield metabolites with the potential to promote neurodegenerative oxidative stress and DNA damage [75]. However, mitochondria also play important roles in other functions that can modulate neurodegeneration, such as apoptosis and endocytosis, and several key AD- and PD-related proteins are localized to mitochondria or the interface between mitochondria and the endoplasmic reticulum [40,76]. The intersection of multiple pathways with mitochondrial function makes this organelle an important target for strategies to combat neurodegeneration.

In AD, the APOE ε4 allele is thought to cause mitochondrial dysfunction through altering the interaction capabilities of its encoded protein’s lipid- and receptor-binding regions [77]. Genes involved in actin pathways, such as CD2AP (CD2-associated protein), may directly impact mitochondrial fission, fusion, and transport along axons due to the dynamic actin remodeling and stabilization required for these processes [78]. Impairment of mitochondrial fission, fusion, and axonal transport can promote abnormal hyperphosphorylation of MAPT (microtubule-associated protein tau), leading to the accumulation of dysfunctional mitochondria and the induction of apoptosis due to poor cellular energetics [79-81]. Components of the mitochondrial membrane are also important for normal functioning through regulation of molecular flux. For example, the AD risk genes TOMM40 (translocase of outer mitochondrial membrane 40 homolog) and TSPO (translocator protein of outer membrane, 18 kDa) are essential for mitochondrial import of proteins and cholesterol, respectively [82,83].

In PD models, SNCA overexpression leads to the excess α-synuclein associating with the mitochondrial membrane and inducing cytochrome c release and oxidative stress [84]. Two other genes associated with both early- and late-onset PD, PARK2 and PINK1 (PTEN-induced putative kinase 1) [85], code for proteins that regulate axonal transport of healthy mitochondria and autophagy of old or dysfunctional mitochondria (also known as mitophagy) [86,87]. Another cause of early-onset PD, PARK7 (parkinson protein 7; also known as DJ1), appears to work in concert with PARK2 and PINK1 as a sensor of oxidative stress and a regulator of mitophagy [84,88].

Interestingly, two compounds used to create experimental models of PD exert their toxic effects in mitochondria. Exposure to MPTP (1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine) was initially proposed as an environmental cause of PD [89]. Since this discovery, injection of MPTP has been used to generate numerous cellular and animal models for PD [90]. In the brain, the MAO-B (monoamine oxidase B) enzyme converts MPTP into MPP+ (1-methyl-4-phenylpyridinium), which interferes with complex I of the mitochondrial electron transport chain to fatally deplete ATP levels and cause neuronal death [90]. The pesticide rotenone is also used to generate PD experimental models and similarly interferes with electron transport chain function [91].

The extensive involvement of mitochondrial stressors and protectors in AD and PD also suggests that changes in mitochondrial DNA might be additional markers of disease. Increased levels of mutations in mitochondrial DNA have been identified in both diseases [92]. However, it is not yet clear whether these mutations affect specific functions or overall mitochondrial health, and it additionally remains to be discovered if specific mitochondrial DNA variants are involved in early-stage disease pathogenesis or if mutations simply provide a measure of ongoing mitochondrial disturbances.

Oxidative DNA damage and repair

Oxidative stress refers to an imbalance between levels of toxic reactive oxygen species (ROS) and the activity of mechanisms – such as the glutathione system and DNA repair pathways – to detoxify ROS to less reactive intermediates or to reverse ROS-induced cellular damage [93]. Mitochondria are the major cellular source of ROS, and therefore dysfunction of mitochondrial components is a significant contributor to oxidative stress and its downstream effects on the structures of DNA, proteins, and lipids. For example, oxidative damage to α-synuclein can change the protein’s targeting sequence to affect its cellular localization and can promote its aggregation [84], and similar mechanisms initiated by oxidative stress have been proposed to affect Aβ as well as other proteins implicated in age-related and neurodegenerative changes [94]. As a result, there is significant interest in whether genetic variation that modulates oxidative stress and its responses might affect susceptibility to neurodegeneration.

The PD-associated genes PARK2, PARK7, and PINK1 may represent one molecular axis contributing to disease risk through regulation of oxidative stress. For example, PARK7 knockdown is known to yield hypersensitivity to oxidative stress in mouse and fly brains [95], while the administration of ROS scavengers and the overexpression of PINK1 and PARK2 have been shown to rescue the effects of PARK7 loss [96]. In AD, disease-associated variants in CLU may inhibit the normal role of clusterin as a protective factor against oxidative stress have been proposed to inhibit the normal role of clusterin as a sensor and chaperone of ROS [97]. Variants in GSTO2 (glutathione S-transferase omega-2), which codes for a subunit of glutathione transferase, have also been associated with decreased levels of glutathione which increase levels of ROS as well as AD susceptibility [98]. Two other genes related to oxidative stress have been identified in large studies of AD-related endophenotypes, including the associations of MTFR1 (mitochondrial fission regulator 1) with cognitive decline [99] and MSRB3 (methionine sulfoxide reductase B3) with hippocampal volume [56].

Oxidative stress and DNA damage repair pathways have also been proposed as points of overlap that might explain the decreased incidence of cancer in individuals with AD or PD other [67,100-102]. It is possible that increased levels of oxidative stress which predispose to neurodegeneration may also harm precancerous cells which would otherwise proceed to unlimited replication. Other mechanisms, such as alternative splicing of genes involved in oxidative metabolism and DNA repair, may also contribute to age- and neurodegenerative disease-associated changes in the brain [103] that oppose the development of cancer. Additional study at the population and molecular levels will be needed to clarify these potential mechanisms.

Ubiquitin-proteasome system

The ubiquitin-proteasome system is responsible for targeted degradation of misfolded, aggregated, or otherwise abnormal proteins. The first step in activating this pathway involves ubiquitin labeling of a protein to direct it to cylindrical proteasomes in the nucleus, endoplasmic reticulum, and other compartments, which recognize ubiquitin-labeled proteins and contain protease enzymes for protein degradation. In contrast to autophagy, which can also degrade proteins in addition to whole organelles,ubiquitin-mediated proteasomal degradation is thought to be highly selective [104].

For AD, PD, and other neurodegenerative diseases marked by accumulation and aggregation of specific abnormal proteins, ubiquitin-proteasome pathways represent natural candidates for modulating pathology. Ubiquitin-positive inclusions in neurons and glial cells are also frequently identified in AD, PD, HD, FTD, and other neurodegenerative disorders and may be a sequelae of dysfunction in proteasomal pathways due to variation in genes including GRN (progranulin) and MAPT among others [105]. Early stages of AD additionally exhibit altered expression of ubiquitin-proteasome pathway genes in astrocytes, which support neuronal function and help maintain homeostasis in the brain [106]. More broadly, ubiquitin-mediated protein degradation may be neuroprotective in modest quantities but may stimulate bulk autophagy or BCL-2-dependent apoptosis at overwhelming or chronic levels [3,107,108].

Interestingly, activation of PD risk genes with direct roles in ubiquitin-proteasome pathways may have beneficial effects in multiple neurodegenerative diseases. For example, UCHL1 (ubiquitin thiolesterase) activation was suggested to reverse AD-associated changes in neuronal dendrite structure through signaling of pathways related to cognition [109,110]. In addition, PARK2 overexpression is proposed to promote clearance of Aβ in AD cell culture models [111], modulate functional levels of SYT11 (synaptotagmin) and other regulators of neurotransmission [112], and increase lifespan and reduce levels of damaged proteins and mitochondria in aging fly brains [113]. These findings corroborate the potential protective effect of ubiquitin-mediated degradation in combating neurodegeneration and highlight the overlapping molecular systems involved in autophagy, mitochondrial regulation, and the ubiquitin-proteasome system.

Local tissue environment

Cell adhesion

Cell adhesion involves the binding of a cell to another cell or to an extracellular surface. In healthy brains, cell adhesion pathways are important for maintenance of synaptic contacts and blood-brain barrier integrity as well as efficient neurotransmission and intracellular signaling [114]. Altered expression of cell adhesion genes is a consistent finding in AD and PD [115-118]. In particular, APOE ε4 may promote neurodegeneration through sequestering targets of RELN (reelin), a protease which signals through APOER2 (apolipoprotein E, receptor 2) and NMDA receptors to enhance synaptic strength and plasticity [119,120]. Depletion of reelin levels in key AD brain regions is thought to be an early event in the disease [121]. Genetic variation in RELN is also associated with AD pathology in cognitively normal older individuals [122], reinforcing the potential role of cell adhesion as an early driver of neurodegenerative changes.

Several studies propose relationships between the Aβ and cell adhesion pathways, including the cleavage of the key synaptic adhesion molecule N-cadherin by presenilin-1 and -2 [123] as well as the interaction of NCAM140 (neuronal cell adhesion molecule 140) with APP to regulate neuronal outgrowth [124]. Recent GWAS of imaging endophenotypes have also identified suggestive associations of cell adhesion genes, including ITGA1 (integrin-α1) and ITGA6 (integrin-α6) with florbetapir positron emission tomography (PET) cerebral Aβ burden [125] and CDH8 (cadherin 8, type 2) with hippocampal volume [126].

In addition, pathway analyses have discovered collective effects on risk among cell adhesion genes in AD and PD. An innovative study integrating PD case-control GWAS and genome-wide expression data for nearly 3500 individuals found enrichment of association for numerous adhesion pathways, including four of the top five results (axon guidance, focal adhesion, cell adhesion molecules, adherens junction) [127]. In AD, cell adhesion pathways have also displayed enrichment of association using case-control GWAS [128] and quantitative trait GWAS of episodic memory impairment [23]. Although cell adhesion genes and pathways are often large and therefore carry risks of false positive associations [19,129], the similarity of findings across these three methodologically diverse studies is striking and provides further support of the hypothesis that adhesion mechanisms can contribute to neurodegeneration.

Endocytosis

The process known as endocytosis, where extracellular molecules are engulfed into membrane-bound vesicles for internalization, is important for gathering nutrients, facilitating molecular interactions and protein degradation in a protected environment, and recycling ligands and receptors [130]. Several AD- and PD-associated genes have central roles in endocytic pathways. For example, APOE is required for microglia to degrade Aβ following endocytosis, and APOE allelic variation affects the efficiency of this degradation in animal models [131]. SORL1 (sortilin 1), whose associations to AD were recently confirmed using exome sequencing [132] and GWAS meta-analysis [133], directs APP to endocytic pathways for recycling and is crucial in preventing the sorting of APP to alternative pathways which generate Aβ [134,135]. In PD, LRRK2 similarly regulates the recycling and/or degradation of α-synuclein [136,137] and is a key influence on the endocytic formation of synaptic vesicles containing neurotransmitters [138].

Promising strategies for therapeutic targeting of endocytic pathways in AD have recently emerged. In an AD mouse model, the retinoid acid receptor (RXR) agonist bexarotene was found to transcriptionally induce APOE to enhance clearance of Aβ and the reversal of cognitive deficits [139]. The yeast homolog of PICALM (phosphatidylinositol binding clathrin assembly protein) is also proposed to be an Aβ toxicity modifier [140]. Thus far, these new findings and their therapeutic implications have not yet been replicated or validated in other systems.

Targeting of endocytic pathways may also be a viable approach to combat PD. The PD-associated gene GAK (cyclin G associated kinase) [59] is a key mediator of endocytic vesicle trafficking by regulating interactions with adaptor proteins and later driving disassembly of the vesicle clathrin coat [141]. In cell culture, under-expression of GAK through knockdown or PD-related mutations accentuates α-synuclein load and toxicity [142]. The closely related gene AAK1 (AP-2 associated kinase 1) [143] has also been associated through GWAS with age of PD onset [144]. The prevalence of disease risk genes and potential drug targets in endocytic pathways is likely to spur continued interest in the coming years.

Neurotransmission

Neurotransmitters are endogenous substances used to relay signals across a synapse. Although overshadowed in recent years by proteinopathy-related theories, initial hypotheses about AD and PD focused on disease-associated neurotransmitter deficits. The selective loss of brain acetylcholine-signaling neurons understood to be crucial for learning and memory drove the hypothesis that AD manifested from a cholinergic deficit [145]. Similarly, the loss of dopaminergic neurons from the substantia nigra understood to be important for motor functioning led to the hypothesis that dysfunction of dopaminergic neurotransmission was a primary cause of PD [146,147]. As a result, modulation of cholinergic or dopaminergic neurotransmission forms the basis of several symptomatic therapies for AD and PD [148,149].

Genetic and molecular studies support a role for neurotransmitter mechanisms in neurodegenerative disease. Pathways related to calcium signaling, which are important for presynaptic neurotransmitter release and postsynaptic signal transduction involving cyclic AMP (cAMP), protein kinase A (PKA), and cAMP response element binding protein (CREB), have displayed association to AD and PD [23,127,128,150,151]. The gene COMT (catechol-O-methyltransferase) encodes an enzyme that degrades dopamine and other catecholamine neurotransmitters, and COMT variants have been associated with dopamine levels in early PD [152] and may contribute to cognitive and psychiatric deficits in AD through interactions with estrogen [153,154]. Further, in addition to its effects on mitochondria, MPP+ gains entry to cells via the dopamine transporter and inhibits synthesis of dopamine and other catecholamines [155,156]. Variation in multiple genes also contributes to elevated glutamate levels in multiple sclerosis (MS), which is classically marked by demyelination and neuroinflammation [157].

Cholinesterase inhibitors, which attempt to increase active levels of acetylcholine in the synaptic cleft by inhibiting the enzymes that degrade acetylcholine, are a first line symptomatic therapy for AD [158]. An initial imaging study in humans identified a correlation between plasma activity of acetylcholinesterase and brain Aβ levels [159]. Recently, a larger study of 555 individuals discovered a genome-wide significant association of variants at the BCHE (butyrylcholinesterase) locus with brain Aβ levels [160]. Butyrylcholinesterase is enriched in senile Aβ plaques [161] and several additional lines of evidence point to potential mechanistic connections among BCHE, APOE, and Aβ [162-165]. Further, some have suggested that cholinesterase inhibitors which preferentially target butyrylcholinesterase may have disease-modifying effects in AD [166,167]. Future work to understand the genetic relationships between the cholinergic and Aβ pathways and their impact on response to drug treatments will be important to improve risk stratification and therapeutic targeting.

Prions and transmissible factors

Prion protein is a membrane-associated, protease-sensitive glycoprotein that is typically enriched in lipid rafts consisting of tightly packed signaling and trafficking molecules [168]. As with other misfolded proteins, misfolded prion protein is normally susceptible to proteasome-mediated and other forms of protein degradation. However, accumulation of misfolded prion protein through inhibition of protein degradation pathways has been proposed to lead to the formation of protease-resistant, aggregated, infectious (i.e., transmissible) particles which can be released to neighboring cells and promote misfolded protein states in those cells [169]. This mechanism is thought to underlie the development of fatal degenerative transmissible spongiform encephalopathies such as Creutzfeld-Jakob disease (CJD), and more controversially has been proposed as a unifying factor promoting neurodegeneration across multiple neurodegenerative diseases including AD, PD, and ALS [170].

So far, genetic association tests of this hypothesis have been mixed, with some studies identifying moderate associations of PRNP (prion protein) variants with neurodegenerative diseases [171-173] and other studies not finding associations [174,175]. Recent GWAS of CJD have also implicated other genes, suggesting that larger pathways related to protein conformational states and prions may be active in neurodegeneration [176-178]. More broadly, a better understanding of the forces contributing to protein conformation and susceptibility to aggregation and transmissibility would be a crucial for unlocking novel diagnostic and therapeutic approaches for neurodegenerative diseases [179]. Genetic variation affecting several related pathways, including translational machinery, endoplasmic reticulum function, chaperone-mediated folding assistance and transportation,and secondary, tertiary, and quaternary protein structural interactions, might represent plausible candidates for association testing to clarify these mechanisms.

Systemic environment

Inflammation and immune dysfunction

Published literature on AD and PD includes robust evidence of disturbances in inflammation and immune pathways. Increased levels of pro-inflammatory cytokines are common findings in blood, cerebrospinal fluid (CSF), and post-mortem brain tissue in both diseases [180-183], and non-steroidal anti-inflammatory drugs have been proposed to have protective effects [184,185]. Active debate has endured on whether inflammation and immune dysregulation are contributors to neurodegeneration or are instead secondary to ongoing cell death. In particular, a fundamental question remains outstanding in neurodegenerative disorders: is inflammation deleterious, protective, or disease stage-dependent?

Studies of microglia, the resident immune system macrophages in the brain and CSF, provide some clues for resolving these issues. Post-mortem tissue analyses as well as newer in vivo PET imaging methods have identified an abundance of activated microglia in AD and PD brains [186]. Both Aβ and α-synuclein are known to activate microglia, stimulating the release of inflammatory cytokines and activation of inflammation-mediating enzymes such as matrix metalloproteinases (MMPs) [186-188]. Activated microglia also express NLRP3 (nucleotide-binding domain and leucine-rich repeat family, pyrin domain containing 3), a component of larger structures known as inflammasomes which promote several inflammatory processes including the maturation of IL-1β (interleukin 1, beta) [189]. In animal models, IL-1β exacerbates AD and PD progression [190,191], and the protective effect of NLRP3 knockout in AD mice likely reflects these underlying mechanisms [192].

Nevertheless, the role of microglia and their secreted products may not be unidirectional. For example, activated microglia are also unique among central nervous system cells in expressing CX3CR1 (chemokine receptor 1), a receptor for the cell survival promoting chemokine known as fractalkine [193]. In PD and ALS mouse models, CX3CR1 knockout resulted in more extensive neuronal loss [194], suggesting that augmentation of signaling through this microglial product may be required for therapy. In addition, microglia may have divergent roles across the course of neurodegenerative diseases. Whereas activation of microglia to stimulate phagocytosis of aggregated disease-related proteins may be protective during early disease stages [195,196], chronic activation of microglia may enhance production of different cytokines which impair phagocytosis and other cell survival-related processes [197].

Genetic associations in inflammation- and immune-related pathways may have similar implications. Variants in IL1B (interleukin 1, beta) and TNFA (tumor necrosis factor, alpha) have been associated with AD and PD and may contribute to altered cytokine levels and inflammatory signaling [198,199]. Meanwhile, AD-associated variants in CLU [97] and TREM2 (triggering receptor expressed on myeloid cells 2) [200-203] may impair the normal anti-inflammatory functions of these genes. TREM2 is predominately expressed on microglia, and recent expression analyses of post-mortem AD human and mouse brain tissue identified perturbations of networks regulated by the TREM2 ligand TYROBP (protein tyrosine kinase binding protein) and enriched with genes functioning in phagocytosis [204], highlighting the potential importance of microglia and their expressed products in modulating neurodegenerative processes.

Other genes appear to bridge inflammation and innate immune responses. For example, the PD- and Crohn’s disease-associated gene LRRK2 both mediates microglial-induced inflammation [205,206] and is a target of IFN-γ (interferon gamma), suggesting an additional role in the immune response to pathogens [207]. Similarly, the AD-associated gene CR1 (complement component receptor 1) [49,208-212] encodes a receptor which may regulate both inflammatory processes as well as classical complement pathways of innate immunity to eliminate synaptic connections [213]. The common involvement of inflammation and immune mechanisms is not limited to AD and PD and appears to extend to ALS [214], MS [27], FTD [215], and psychiatric disorders [216].

These findings suggest that fulfilling the promise of therapies targeting these pathways in neurodegenerative disease might be quite complex [183,217-219]. Appropriate modulation of inflammatory and immune mechanisms may require combinatorial regulation of multiple factors, with some being activated and others deactivated depending on disease stage and an individual’s genetic profile.

Lipid, metabolic, and endocrine factors

Recent epidemiological and molecular studies are converging to support the hypothesis that loss of lipid homeostasis can prominently contribute to neurodegeneration. Findings that atherosclerosis and other cardiovascular diseases are impacted by APOE ε4 and can increase the risk of AD [220] are complemented by studies suggesting that statin use to lower circulating cholesterol may modestly reduce the risk of AD and PD [221,222]. Importantly, neuronal membranes contain substantial amounts of cholesterol and other lipids, and disturbances in lipid pathways have been frequently proposed to impact synaptic signaling and neuronal plasticity and degeneration [223-226].

As the major lipoprotein of the brain, ApoE transports key lipids and associated proteins to cells for uptake via receptor-mediated endocytosis [220]. The degree of lipidation in ApoE is an important factor in maintaining lipid homeostasis and in mediating interactions with Aβ which can promote its endocytic clearance, and APOE allelic variants may affect both processes [227]. Strikingly, two other AD GWAS-implicated genes have primary roles in lipid homeostasis: CLU represents the second major lipoprotein of the brain (also known as apolipoprotein J) [6,228] and ABCA7 codes for a microglia-enriched trans-membrane cholesterol and phospholipid transporter [229,230]. Among PD-related genes, both PARK2 and LRRK2 code for proteins which regulate cellular uptake of lipid-rich structures [231-233].

Recently, lipidomics analyses of the complete profile of lipids and their metabolites in tissue samples have provided initial unbiased views of lipid pathway disturbances in AD and PD [225,234,235]. In PD, this approach identified changes in lipid metabolism in human primary visual cortex, a region that does not exhibit significant Lewy body pathology but may be important for visual symptoms in PD [235]. These large-scale findings reinforce the concept that lipid pathways are highly complex and include numerous components with the potential for local and remote impacts on inflammation, oxidative stress, vascular, and other pathways. As a result, drugs targeting lipid pathways, including supplementary administration of endogenous compounds [236], would be expected to have pleiotropic effects in the context of neurodegenerative disease which may require modulation based on the functional status of other pathways in an individual [237].

Among metabolic disorders, a particularly interesting relationship is apparent between diabetes and AD. The presence of type 2 diabetes doubles the risk of AD [238] and metabolic dysregulation, including loss of insulin signaling through the PI3 kinase and AKT, occurs in the brain in early AD [239]. In addition, models of insulin resistance or deficiency result in cerebral Aβ buildup while models of Aβ toxicity lead to decreased insulin signaling [240]. As a result, diabetes and AD may share several drug targets, including insulin and IGF (insulin-like growth factor) stimulation [241,242], inflammation [243], BCHE [160,244,245], and GSK3 (glycogen synthase kinase 3) [246].

Vascular changes

Vascular pathology, including increases in vessel wall stiffness, changes in endothelial cell adhesion and metabolism, and dysfunction of the blood-brain barrier, can promote neurodegeneration through yielding chronic, low perfusion [247]. Presence of the APOE ε4 allele is a well-known risk factor for dyslipidemia, atherosclerosis, and coronary heart disease [100,248], suggesting that part of the impact of APOE on AD may be mediated through vascular mechanisms. Pathological changes to the blood-brain barrier have also been identified in AD and PD through histological and molecular analyses and may explain the proposed modest protective effect of caffeine intake in these diseases [249-251].

Vascular smooth muscle pathways have displayed genetic associations with AD imaging phenotypes [252], and the AD-associated gene CR1 was also found to increase the risk of cerebral amyloid angiopathy, a leading cause of intracerebral hemorrhage in older individuals [253]. Although other vascular-related genes such as VEGF (vascular endothelial growth factor) have displayed mostly mixed results in association tests for AD and PD [254,255], additional studies will be important to determine the effects of in situ genetic risk factors on vascular functioning and brain plasticity [256], relationships of vascular pathways to other mechanisms of neurodegeneration [257], and the impact of lifestyle measures such as healthy diet and exercise on disease onset and progression. Comparisons of genetic and environmental risk factors for AD and PD with those impacting vascular dementia will also illuminate common and discordant features of their underlying pathophysiology [258].

Neurodevelopment and biological aging

Epigenetic changes

Epigenetic factors provide mechanisms for genetic control that do not involve modifications to an individual’s DNA sequence [259]. These heritable changes, including RNA-associated silencing and methylation or acetylation of DNA or histones, can dynamically respond to environmental stimuli [260] and also appear to increase in frequency with aging [261]. Several AD- and PD-related genes are regulators or targets of epigenetic mechanisms. For example, nuclear α-synuclein accumulation inhibits histone acetylation and promotes apoptosis in cell culture [262]. While PD-related SNCA mutations potentiate this effect, inhibition of SIRT2 (sirtuin 2) deacetylase activity may reverse SNCA-induced toxicity [263]. Similarly, inhibition of HDAC2 (histone deacetylase 2) facilitates expression of genes related to learning and memory and reverses AD symptoms in mice [264]. Epigenetic pathways may also impact Aβ pathology: in mice, SIRT1 (sirtuin 1) deacetylase activity promotes the alternative cleavage of APP by ADAM10 (α-secretase) to decrease formation of Aβ [265]. In addition, nucleotide repeat expansions in C9orf72 (chromosome 9 open reading frame 72), which are a major cause of familial FTD, ALS, and related neurodegenerative disorders [266], may exert their pathologic effects via mechanisms related to RNA-mediated silencing or unconventional translation [267,268].

Human epigenome-wide studies have not yet been reported for AD or PD. In analyses of candidate genes related to neuroinflammation and synaptic functioning, changes in methylation of CpG islands in the promoters of BDNF (brain-derived neurotrophic factor), COX2 (cyclooxygenase-2), CREB (cyclic AMP response element binding protein), and NFKB (nuclear factor kappa B) were identified in human post-mortem AD frontal cortex [269]. Epigenome-wide studies might discover novel loci contributing to AD and PD and would be particularly informative for early stages of the disease spectrum, where targeted therapies would likely be most effective, and to capture dynamic changes in epigenetic markers longitudinally.

Neurotrophic factors

Neurotrophic factors (neurotrophins) are secreted growth factors that promote the development, functioning, and survival of neurons through regulation of gene transcription. Neurotrophins typically affect transcription through binding receptors at neuron terminals to stimulate second messenger signaling cascades or to promote their internalization and direct transport along the axon to the nucleus [270]. Diminished signaling and axonal transport of BDNF and NGF (nerve growth factor) have been identified in post-mortem AD brain tissue [271], and variants in BDNF have been associated with CSF Aβ levels in AD [272], age of onset in familial PD [273], and age-related changes in brain structure and cognitive function in individuals without frank disease [274], suggesting a primary role for neurotrophin signaling in susceptibility to neurodegeneration.

Novel treatment approaches for augmenting neurotrophin signaling appear promising for enhancing neuronal survival and functioning to combat degenerative changes. For example, exogenous administration of BDNF was observed to rescue stress hormone-induced AD-like memory impairment in rats through activation of several memory-related signaling pathways [275]. In addition, SNPs in the dopaminergic neurotrophin gene CDNF (cerebral dopamine neurotrophic factor) have been associated with PD risk [276], and the highly related gene GDNF (glial cell derived neurotrophic factor) is also being explored as a potential therapeutic target for PD [277,278]. It should be noted that neurotrophins can be expressed in non-neuronal tissue and may have roles in promoting or inhibiting cancer at those sites [279,280] which will require further evaluation in the context of potential neurotrophin-related treatment strategies for neurodegenerative disease.

Telomeres

Telomeres are DNA sequences at the ends of chromosomes that provide protection against the loss of more proximal genetic material during DNA replication in mitosis [281]. In germ-line and some somatic cells, the enzyme telomerase is responsible for maintaining telomere length and structure. However, most adult somatic cells do not express telomerase and as a result gradually lose telomere length and structure with each cycle of mitosis. While reactivation of telomerase contributes to many types of cancer by maintaining a limitless proliferative ability for tumor cells, excessively short telomere length in aging cells is proposed to signal for senescence and apoptosis [281,282].

Although shortened telomere length in peripheral white blood cells has been associated with dementia and mortality in older adults, even after adjusting for APOE genotype [283], the relationship between telomere length in neurons and neurodegeneration is not yet clear. In one study, neuronal telomere shortening induced microglial proliferation (microgliosis) in aging mice but reduced microgliosis and Aβ pathology while improving memory and learning in AD mice [284]. Changes in telomere length have not been widely observed in peripheral white blood cells or in the brain in PD or ALS but will likely receive continued scrutiny [285-287]. In particular, several genetic influences on telomere length have been identified which may provide novel candidates for study in relation to neurodegenerative disease. Variants in TERC (telomerase RNA component), which codes for a component of telomerase, have been associated with telomere length in several human study samples [288,289], as have genes related to DNA and histone methylation [290]. In addition, telomere pathways have exhibited enrichment of genetic association to human longevity in a large cohort study [291]. These preliminary findings suggest that neurodegenerative diseases may be amenable to therapies targeting mechanisms of cellular and biological aging more broadly [282,292].

Network analysis of top AD- and PD-associated genes

To complement the pathway-driven approach, we performed network analysis to identify additional functional relationships between top AD- and PD-associated genes. While pathways are defined by overarching goals and the mechanistic steps involved, networks can display other types of relationships which may cut across multiple pathways or may indicate novel pathways which have not yet been characterized [19].

Due to the numerous pathways implicated in AD and PD and the pleiotropic effects of many key disease-associated genes, we hypothesized that regulatory relationships among these genes might impact multiple pathways. To explore this hypothesis, we performed transcription factor network analysis using the MetaCore software (GeneGo, Inc.). This approach incorporates knowledge from published literature to relate an input list of genes to known transcription factors and proximal targets such as ligand-receptor interactions. As input, we used the top 10 genes from the AlzGene (APOE, BIN1, CLU, ABCA7, CR1, PICALM, MS4A6A, CD33, MS4A4E, CD2AP) [293] and PDGene (MAPT, SNCA, GBA, LRRK2, PM20D1, GAK, MCCC1, STK39, BST1, GPNMB) [294] databases in addition to a small number of genes (APP, PSEN1, PSEN2, DJ1, HIP1R, PARK2, SYT11, UCHL) implicated in both Mendelian and sporadic forms of AD or PD.

A network was identified which displays relationships among 31 factors, including 19 of the 28 input genes (Figure 2). The probability of the software algorithm generating a network with this level of interconnectedness by random selection of input genes was exceedingly small (p = 1.14 × 10-54). Strikingly, numerous genes in the network exhibit co-regulation by the SP1 (specificity protein 1) and AP-1 (activating protein 1) transcription factors. SP1 has been previously noted to regulate the expression of multiple AD-related genes [23,295]. Elevated levels of SP1 have been identified in AD human brains and mouse models [296,297] and may be induced by inflammation and oxidative stress [296,298]. The AP-1 transcription factor is composed of heterodimers of several proteins, including those encoded by the FOS and JUN proto-oncogenes [299]. AP-1 is an important regulator of dopaminergic signaling pathways [300,301] as well as numerous genes related to autophagy and lysosomal function [302]. Interestingly, animal models indicate that inhibition of SP1 may be neuroprotective in AD [297] and inhibition of AP-1 may be neuroprotective in PD [303]. The connections among SP1, AP-1, and AD- and PD-associated genes suggest that coordinate modulation of these transcription factors may be a viable strategy for combating neurodegeneration.

Figure 2.

Figure 2

Regulatory network centered on the SP1 and AP-1 transcription factors is enriched with top AD and PD genes. Meta-analytic genetic association data from public databases and supplementary manual curation was used to generate a list of 13 AD genes and 15 PD genes. Network analysis was performed using MetaCore (GeneGo, Inc.) to relate these input genes to known transcription factors and proximal targets based on published findings. A highly interconnected network including 9 AD genes (labeled in blue), 10 PD genes (labeled in red), and 13 additional genes (labeled in black) was identified. Many of the input AD and PD genes exhibit co-regulation by the SP1 and AP-1 transcription factors. Other genes of interest were also related to input AD and PD genes and represent a variety of candidate pathways in neurodegeneration.

This transcriptional network also includes several additional genes of interest which were not in the initial input list. For example, EGR1 (early growth response 1) encodes a zinc-finger transcription factor that is important for synaptic plasticity [304] and cognitive performance [305] and whose up-regulation in AD brains may promote phosphorylation of tau [306]. The transcription factor encoded by HMGB1 (high-mobility group protein 1) can also directly bind aggregated α-synuclein [307], regulate phagocytosis of Aβ [308,309], and promote inflammation when secreted by activated microglia or necrotic neurons [310,311]. Interactions between HIV-1 TAT (transactivator of transcription) and genes involved in AD and PD may be involved in HIV-associated cognitive impairment and Aβ pathology [312,313]. Other genes of interest in this regulatory network include MMP9 (matrix metalloproteinase 9) which is involved in synaptic plasticity and Aβ degradation [314], IRF3 and IRF7 (interferon regulatory factors 3 and 7) which regulate interferon-mediated inflammation and immune responses [315-318], and LRP1 (low density lipoprotein receptor-related protein 1) which may affect several neurodegeneration pathways including lipid metabolism, Aβ endocytosis, and inflammation [319-322].

It should be noted that this analysis is not comprehensive or unbiased. Complementary strategies, including the use of alternative criteria for selection or statistical weighting of input genes as well as other schema for defining network connections, might highlight different relationships. Nevertheless, this regulatory network generates hypotheses for further investigation and reflects, at the transcriptional level, many of the same pathways implicated by genetic studies of AD and PD. More broadly, a better understanding of altered transcriptional regulation patterns through whole genome expression arrays and whole transcriptome sequencing (RNA-seq) would augment GWAS findings in neurodegenerative disease and would provide functional information to connect genetic associations with their biochemical outcomes.

Conclusions and future prospects

Through a detailed review of GWAS, we identified numerous pathways common to AD and PD which nominate promising new targets for further study as well as biomarker and drug development. These findings build on established notions of complex disease etiology, with multiple processes presumed to influence neurodegeneration and clinical outcomes in AD, PD, and related disorders. They also advance the understanding of mechanisms likely to be crucial in maintaining brain structure and function during normal aging, in contrast to changes seen in AD and PD. These insights suggest that collaborative efforts to leverage genetic and biomarker data in AD, PD, and related disorders would likely provide major stimuli for developing unified treatment approaches to combat neurodegeneration.

For neurodegenerative and other complex diseases, accounting for the substantial heritability that is not explained by individual GWAS-implicated variants is an ongoing challenge [11,323]. The pathways and networks identified here provide several routes for addressing this limitation. For example, pathway analysis of GWAS data relies on high quality pathway definitions, and for some biological realms, expert and updated manual curation of pathways can be superior to public databases and enhance statistical power for these analyses [19,20]. Pathways implicated by common SNPs from GWAS also provide a knowledge-driven framework for targeting initial studies with WGS data, which is better suited for detection of rare SNPs and copy number and other structural variants but is computationally demanding to store and analyze [31]. Finally, interactions among known variants and lifestyle, environmental, and epigenetic factors may impact susceptibility [324], and pathways and networks understood to be involved in pathogenesis may be more likely to contain these interactions [36,38].

Diagnosis and treatment strategies for neurodegenerative diseases may also need to evolve to reflect a complex genetic architecture involving multiple pathways. One possibility is that a combination of clinical biomarkers – such as genotype, blood and CSF analyte, brain imaging, cognitive assessment, and medical history data – might be required in order to detect the effects of multiple pathways. Since the functions of many disease pathways may be disease stage-specific, high blood levels of a particular cytokine might have different implications for risk stratification depending on genotype, brain structure, and other measures. Similarly, therapeutic and preventative strategies for neurodegenerative disease may benefit from drug combinations based on the cocktail approaches used for HIV infection and some cancers. It is possible that efficacy, and therefore the choice of particular drugs to include in the cocktail, may depend on an individual’s profile of biomarkers and key genetic variants – some of which may be protective and others deleterious – in targeted pathways. The development of advanced statistical models for analysis of large, multimodal datasets will help to explore these potentially new paradigms that may facilitate a personalized medicine for neurodegenerative diseases.

More broadly, pathways and networks can serve as vehicles for integrating findings from diverse studies of neurodegeneration. There are many active strategies for large scale omics analysis of neurodegenerative disease (Figure 3), and findings that converge across these multiple study designs can provide confirmatory evidence that is crucial for efficient clinical translation. Isolated genes and molecules can be challenging to evaluate for convergence since they may not be represented in all data modalities or experimental model systems. In contrast, pathways and networks can incorporate data from multiple biological levels (e.g., genes, transcripts, proteins, and metabolites, among others) and may be more likely to be evolutionarily conserved [325]. For example, recent pathway-based studies integrating GWAS and gene expression data have demonstrated enhanced power, reproducibility, and connections of top findings to hypothesized disease processes [127,326,327]. The utility of these studies will increase as present limitations of pathway-based approaches are addressed, including how to incorporate associations from intergenic regions and from genes without known functions. A pathway-based framework also emphasizes that the discovery of a strongly associated genetic variant represents a foundation to study functionally related genes, since other components in the pathway may yield better targets for biomarker and drug development [29,30,328]. These advantages will be vital in harnessing the wealth of existing data on neurodegenerative disease to develop an integrated understanding of its mechanisms and formulate optimal clinical guidelines.

Figure 3.

Figure 3

Biological pathways and networks: a hub for convergent omics. Numerous large scale omics approaches are being used to study complex neurodegenerative diseases and endophenotypes in human tissue and animal and other model systems. Unlike individual genes and other isolated molecules, which may not be present in all model systems and may have differential sensitivity for detection with various study designs, pathways and networks are well-conserved and can be evaluated for convergence across diverse methodological approaches. Integration of findings to identify pathways and networks with consistent relationships to disease is likely to enhance the development of diagnostic biomarkers and treatment and prevention strategies.

Acknowledgements

This work was supported by the Indiana Clinical and Translational Sciences Institute (CTSI) National Institutes of Health (NIH) grants U54 RR025761, RR027710-01, and RR020128, the Indiana University Medical Scientist Training Program (MSTP) NIH grant GM077229-02, the National Science Foundation (NSF) grant IIS-11173365, as well as NIH grants U01 AG024904, RC2 AG036535, R01 AG19771, P30 AG10133, and R01 LM011360.

Disclosure of conflict of interest

The authors declare no conflict of interest.

References

  • 1.Hardy JA, Higgins GA. Alzheimer’s disease: the amyloid cascade hypothesis. Science. 1992;256:184–185. doi: 10.1126/science.1566067. [DOI] [PubMed] [Google Scholar]
  • 2.Karran E, Mercken M, De Strooper B. The amyloid cascade hypothesis for Alzheimer’s disease: an appraisal for the development of therapeutics. Nat Rev Drug Discov. 2011;10:698–712. doi: 10.1038/nrd3505. [DOI] [PubMed] [Google Scholar]
  • 3.Taylor JP, Hardy J, Fischbeck KH. Toxic proteins in neurodegenerative disease. Science. 2002;296:1991–1995. doi: 10.1126/science.1067122. [DOI] [PubMed] [Google Scholar]
  • 4.Krainc D. Clearance of mutant proteins as a therapeutic target in neurodegenerative diseases. Arch Neurol. 2010;67:388–392. doi: 10.1001/archneurol.2010.40. [DOI] [PubMed] [Google Scholar]
  • 5.Rademakers R, Neumann M, Mackenzie IR. Advances in understanding the molecular basis of frontotemporal dementia. Nat Rev Neurol. 2012;8:423–434. doi: 10.1038/nrneurol.2012.117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Sleegers K, Lambert JC, Bertram L, Cruts M, Amouyel P, Van Broeckhoven C. The pursuit of susceptibility genes for Alzheimer’s disease: progress and prospects. Trends Genet. 2010;26:84–93. doi: 10.1016/j.tig.2009.12.004. [DOI] [PubMed] [Google Scholar]
  • 7.Hardy J. Genetic analysis of pathways to Parkinson disease. Neuron. 2010;68:201–206. doi: 10.1016/j.neuron.2010.10.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Gandhi S, Wood NW. Genome-wide association studies: the key to unlocking neurodegeneration? Nat Neurosci. 2010;13:789–794. doi: 10.1038/nn.2584. [DOI] [PubMed] [Google Scholar]
  • 9.Bertram L, Tanzi RE. The genetic epidemiology of neurodegenerative disease. J Clin Invest. 2005;115:1449–1457. doi: 10.1172/JCI24761. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.McCarthy MI, Abecasis GR, Cardon LR, Goldstein DB, Little J, Ioannidis JP, Hirschhorn JN. Genome-wide association studies for complex traits: consensus, uncertainty and challenges. Nat Rev Genet. 2008;9:356–369. doi: 10.1038/nrg2344. [DOI] [PubMed] [Google Scholar]
  • 11.Chee Seng K, En Yun L, Yudi P, Kee Seng C. The pursuit of genome-wide association studies: where are we now? J Hum Genet. 2010;55:195–206. doi: 10.1038/jhg.2010.19. [DOI] [PubMed] [Google Scholar]
  • 12.Noorbakhsh F, Overall CM, Power C. Deciphering complex mechanisms in neurodegenerative diseases: the advent of systems biology. Trends Neurosci. 2009;32:88–100. doi: 10.1016/j.tins.2008.10.003. [DOI] [PubMed] [Google Scholar]
  • 13.Hindorff LA, Junkins HA, Hall PN, Mehta JP, Manolio TA. A Catalog of Published Genome-Wide Association Studies. 2011. National Human Genome Research Institute, http://www.genome.gov/gwastudies.
  • 14.Gatz M, Reynolds CA, Fratiglioni L, Johansson B, Mortimer JA, Berg S, Fiske A, Pedersen NL. Role of genes and environments for explaining Alzheimer disease. Arch Gen Psychiatry. 2006;63:168–174. doi: 10.1001/archpsyc.63.2.168. [DOI] [PubMed] [Google Scholar]
  • 15.Kamboh MI, Demirci FY, Wang X, Minster RL, Carrasquillo MM, Pankratz VS, Younkin SG, Saykin AJ, Jun G, Baldwin C, Logue MW, Buros J, Farrer L, Pericak-Vance MA, Haines JL, Sweet RA, Ganguli M, Feingold E, DeKosky ST, Lopez OL, Barmada MM. Genome-wide association study of Alzheimer’s disease. Transl Psychiatry. 2012;2:e117. doi: 10.1038/tp.2012.45. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Cross-Disorder Group of the Psychiatric Genomics Consortium. Smoller JW, Craddock N, Kendler K, Lee PH, Neale BM, Nurnberger JI, Ripke S, Santangelo S, Sullivan PF. Identification of risk loci with shared effects on five major psychiatric disorders: a genome-wide analysis. Lancet. 2013 Apr 20;381:1371–9. doi: 10.1016/S0140-6736(12)62129-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Simon-Sanchez J, Singleton A. Genome-wide association studies in neurological disorders. Lancet Neurol. 2008;7:1067–1072. doi: 10.1016/S1474-4422(08)70241-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Cantor RM, Lange K, Sinsheimer JS. Prioritizing GWAS results: A review of statistical methods and recommendations for their application. Am J Hum Genet. 2010;86:6–22. doi: 10.1016/j.ajhg.2009.11.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Ramanan VK, Shen L, Moore JH, Saykin AJ. Pathway analysis of genomic data: concepts, methods, and prospects for future development. Trends Genet. 2012;28:323–332. doi: 10.1016/j.tig.2012.03.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Wang K, Li M, Hakonarson H. Analysing biological pathways in genome-wide association studies. Nat Rev Genet. 2010;11:843–854. doi: 10.1038/nrg2884. [DOI] [PubMed] [Google Scholar]
  • 21.Lambert JC, Grenier-Boley B, Chouraki V, Heath S, Zelenika D, Fievet N, Hannequin D, Pasquier F, Hanon O, Brice A, Epelbaum J, Berr C, Dartigues JF, Tzourio C, Campion D, Lathrop M, Amouyel P. Implication of the immune system in Alzheimer’s disease: evidence from genome-wide pathway analysis. J Alzheimers Dis. 2010;20:1107–1118. doi: 10.3233/JAD-2010-100018. [DOI] [PubMed] [Google Scholar]
  • 22.Hong MG, Alexeyenko A, Lambert JC, Amouyel P, Prince JA. Genome-wide pathway analysis implicates intracellular transmembrane protein transport in Alzheimer disease. J Hum Genet. 2010;55:707–709. doi: 10.1038/jhg.2010.92. [DOI] [PubMed] [Google Scholar]
  • 23.Ramanan VK, Kim S, Holohan K, Shen L, Nho K, Risacher SL, Foroud TM, Mukherjee S, Crane PK, Aisen PS, Petersen RC, Weiner MW, Saykin AJ; Alzheimer’s Disease Neuroimaging Initiative (ADNI) 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;6:634–648. doi: 10.1007/s11682-012-9196-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.O’Dushlaine C, Kenny E, Heron E, Donohoe G, Gill M, Morris D, Corvin A. Molecular pathways involved in neuronal cell adhesion and membrane scaffolding contribute to schizophrenia and bipolar disorder susceptibility. Mol Psychiatry. 2011;16:286–292. doi: 10.1038/mp.2010.7. [DOI] [PubMed] [Google Scholar]
  • 25.O’Dushlaine C, Kenny E, Heron EA, Segurado R, Gill M, Morris DW, Corvin A. The SNP ratio test: pathway analysis of genome-wide association datasets. Bioinformatics. 2009;25:2762–2763. doi: 10.1093/bioinformatics/btp448. [DOI] [PubMed] [Google Scholar]
  • 26.Baranzini SE, Galwey NW, Wang J, Khankhanian P, Lindberg R, Pelletier D, Wu W, Uitdehaag BM, Kappos L, Polman CH, Matthews PM, Hauser SL, Gibson RA, Oksenberg JR, Barnes MR. Pathway and network-based analysis of genome-wide association studies in multiple sclerosis. Hum Mol Genet. 2009;18:2078–2090. doi: 10.1093/hmg/ddp120. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Sawcer S, Hellenthal G, Pirinen M, Spencer C, Patsopoulos N, Moutsianas L, Dilthey A, Su Z, Freeman C, Hunt S, Edkins S, Gray E, Booth D, Potter SC, Goris A, Band G, Oturai AB, Strange A, Saarela J, Bellenguez C, Fontaine B, Gillman M, Hemmer B, Gwilliam R, Zipp F, Jayakumar A, Martin R, Leslie S, Hawkins S, Giannoulatou E, D’alfonso S, Blackburn H, Boneschi FM, Liddle J, Harbo HF, Perez ML, Spurkland A, Waller MJ, Mycko MP, Ricketts M, Comabella M, Hammond N, Kockum I, McCann OT, Ban M, Whittaker P, Kemppinen A, Weston P, Hawkins C, Widaa S, Zajicek J, Dronov S, Robertson N, Bumpstead SJ, Barcellos LF, Ravindrarajah R, Abraham R, Alfredsson L, Ardlie K, Aubin C, Baker A, Baker K, Baranzini SE, Bergamaschi L, Bergamaschi R, Bernstein A, Berthele A, Boggild M, Bradfield JP, Brassat D, Broadley SA, Buck D, Butzkueven H, Capra R, Carroll WM, Cavalla P, Celius EG, Cepok S, Chiavacci R, Clerget-Darpoux F, Clysters K, Comi G, Cossburn M, Cournu-Rebeix I, Cox MB, Cozen W, Cree BA, Cross AH, Cusi D, Daly MJ, Davis E, de Bakker PI, Debouverie M, D’hooghe MB, Dixon K, Dobosi R, Dubois B, Ellinghaus D, Elovaara I, Esposito F, Fontenille C, Foote S, Franke A, Galimberti D, Ghezzi A, Glessner J, Gomez R, Gout O, Graham C, Grant SF, Guerini FR, Hakonarson H, Hall P, Hamsten A, Hartung HP, Heard RN, Heath S, Hobart J, Hoshi M, Infante-Duarte C, Ingram G, Ingram W, Islam T, Jagodic M, Kabesch M, Kermode AG, Kilpatrick TJ, Kim C, Klopp N, Koivisto K, Larsson M, Lathrop M, Lechner-Scott JS, Leone MA, Leppä V, Liljedahl U, Bomfim IL, Lincoln RR, Link J, Liu J, Lorentzen AR, Lupoli S, Macciardi F, Mack T, Marriott M, Martinelli V, Mason D, McCauley JL, Mentch F, Mero IL, Mihalova T, Montalban X, Mottershead J, Myhr KM, Naldi P, Ollier W, Page A, Palotie A, Pelletier J, Piccio L, Pickersgill T, Piehl F, Pobywajlo S, Quach HL, Ramsay PP, Reunanen M, Reynolds R, Rioux JD, Rodegher M, Roesner S, Rubio JP, Rückert IM, Salvetti M, Salvi E, Santaniello A, Schaefer CA, Schreiber S, Schulze C, Scott RJ, Sellebjerg F, Selmaj KW, Sexton D, Shen L, Simms-Acuna B, Skidmore S, Sleiman PM, Smestad C, Sørensen PS, Søndergaard HB, Stankovich J, Strange RC, Sulonen AM, Sundqvist E, Syvänen AC, Taddeo F, Taylor B, Blackwell JM, Tienari P, Bramon E, Tourbah A, Brown MA, Tronczynska E, Casas JP, Tubridy N, Corvin A, Vickery J, Jankowski J, Villoslada P, Markus HS, Wang K, Mathew CG, Wason J, Palmer CN, Wichmann HE, Plomin R, Willoughby E, Rautanen A, Winkelmann J, Wittig M, Trembath RC, Yaouanq J, Viswanathan AC, Zhang H, Wood NW, Zuvich R, Deloukas P, Langford C, Duncanson A, Oksenberg JR, Pericak- Vance MA, Haines JL, Olsson T, Hillert J, Ivinson AJ, De Jager PL, Peltonen L, Stewart GJ, Hafler DA, Hauser SL, McVean G, Donnelly P, Compston A. Genetic risk and a primary role for cell-mediated immune mechanisms in multiple sclerosis. Nature. 2011;476:214–219. doi: 10.1038/nature10251. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Psychiatric GWAS Consortium Bipolar Disorder Working Group. Large-scale genome-wide association analysis of bipolar disorder identifies a new susceptibility locus near ODZ4. Nat Genet. 2011;43:977–983. doi: 10.1038/ng.943. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Hirschhorn JN. Genomewide Association Studies — Illuminating Biologic Pathways. N Engl J Med. 2009;360:1699–1701. doi: 10.1056/NEJMp0808934. [DOI] [PubMed] [Google Scholar]
  • 30.Penrod NM, Cowper-Sal-lari R, Moore JH. Systems genetics for drug target discovery. Trends Pharmacol Sci. 2011;32:623–630. doi: 10.1016/j.tips.2011.07.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Bras J, Guerreiro R, Hardy J. Use of next-generation sequencing and other whole-genome strategies to dissect neurological disease. Nat Rev Neurosci. 2012;13:453–464. doi: 10.1038/nrn3271. [DOI] [PubMed] [Google Scholar]
  • 32.Braskie MN, Ringman JM, Thompson PM. Neuroimaging measures as endophenotypes in Alzheimer’s disease. Int J Alzheimers Dis. 2011;2011:490140. doi: 10.4061/2011/490140. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Kendler KS, Neale MC. Endophenotype: a conceptual analysis. Mol Psychiatry. 2010;15:789–797. doi: 10.1038/mp.2010.8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Cruchaga C, Kauwe JS, Nowotny P, Bales K, Pickering EH, Mayo K, Bertelsen S, Hinrichs A Alzheimer’s Disease Neuroimaging Initiative. Fagan AM, Holtzman DM, Morris JC, Goate AM. Cerebrospinal fluid APOE levels: an endophenotype for genetic studies for Alzheimer’s disease. Hum Mol Genet. 2012 Oct 15;21:4558–71. doi: 10.1093/hmg/dds296. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Jack CR Jr, Vemuri P, Wiste HJ, Weigand SD, Lesnick TG, Lowe V, Kantarci K, Bernstein MA, Senjem ML, Gunter JL, Boeve BF, Trojanowski JQ, Shaw LM, Aisen PS, Weiner MW, Petersen RC, Knopman DS Alzheimer’s Disease Neuroimaging Initiative. Shapes of the Trajectories of 5 Major Biomarkers of Alzheimer Disease. Arch Neurol. 2012;69:856–867. doi: 10.1001/archneurol.2011.3405. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.McKinney BA, Pajewski NM. Six Degrees of Epistasis: Statistical Network Models for GWAS. Front Genet. 2011;2:109. doi: 10.3389/fgene.2011.00109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Vidal M, Cusick ME, Barabasi AL. Interactome networks and human disease. Cell. 2011;144:986–998. doi: 10.1016/j.cell.2011.02.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Schadt EE. Molecular networks as sensors and drivers of common human diseases. Nature. 2009;461:218–223. doi: 10.1038/nature08454. [DOI] [PubMed] [Google Scholar]
  • 39.Nussbaum RL, Ellis CE. Alzheimer’s Disease and Parkinson’s Disease. N Engl J Med. 2003;348:1356–1364. doi: 10.1056/NEJM2003ra020003. [DOI] [PubMed] [Google Scholar]
  • 40.Bredesen DE, Rao RV, Mehlen P. Cell death in the nervous system. Nature. 2006;443:796–802. doi: 10.1038/nature05293. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Mattson MP. Apoptosis in neurodegenerative disorders. Nat Rev Mol Cell Biol. 2000;1:120–129. doi: 10.1038/35040009. [DOI] [PubMed] [Google Scholar]
  • 42.Loo DT, Copani A, Pike CJ, Whittemore ER, Walencewicz AJ, Cotman CW. Apoptosis Is Induced by Beta-Amyloid in Cultured Central-Nervous-System Neurons. Proc Natl Acad Sci U S A. 1993;90:7951–7955. doi: 10.1073/pnas.90.17.7951. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Wolozin B, Iwasaki K, Vito P, Ganjei JK, Lacana E, Sunderland T, Zhao B, Kusiak JW, Wasco W, D’Adamio L. Participation of presenilin 2 in apoptosis: enhanced basal activity conferred by an Alzheimer mutation. Science. 1996;274:1710–1713. doi: 10.1126/science.274.5293.1710. [DOI] [PubMed] [Google Scholar]
  • 44.Weidemann A, Paliga K, Durrwang U, Reinhard FBM, Schuckert O, Evin G, Masters CL. Proteolytic processing of the Alzheimer’s disease amyloid precursor protein within its cytoplasmic domain by caspase-like proteases. J Biol Chem. 1999;274:5823–5829. doi: 10.1074/jbc.274.9.5823. [DOI] [PubMed] [Google Scholar]
  • 45.Guo Q, Sebastian L, Sopher BL, Miller MW, Glazner GW, Ware CB, Martin GM, Mattson MP. Neurotrophic factors [activity-dependent neurotrophic factor (ADNF) and basic fibroblast growth factor (bFGF)] interrupt excitotoxic neurodegenerative cascades promoted by a PS1 mutation. Proc Natl Acad Sci U S A. 1999;96:4125–4130. doi: 10.1073/pnas.96.7.4125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.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;261:921–923. doi: 10.1126/science.8346443. [DOI] [PubMed] [Google Scholar]
  • 47.Koffie RM, Hashimoto T, Tai HC, Kay KR, Serrano-Pozo A, Joyner D, Hou S, Kopeikina KJ, Frosch MP, Lee VM, Holtzman DM, Hyman BT, Spires-Jones TL. Apolipoprotein E4 effects in Alzheimer’s disease are mediated by synaptotoxic oligomeric amyloid-beta. Brain. 2012;135:2155–2168. doi: 10.1093/brain/aws127. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Soler-Lopez M, Zanzoni A, Lluis R, Stelzl U, Aloy P. Interactome mapping suggests new mechanistic details underlying Alzheimer’s disease. Genome Res. 2011;21:364–376. doi: 10.1101/gr.114280.110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Lambert JC, Heath S, Even G, Campion D, Sleegers K, Hiltunen M, Combarros O, Zelenika D, Bullido MJ, Tavernier B, Letenneur L, Bettens K, Berr C, Pasquier F, Fievet N, Barberger-Gateau P, Engelborghs S, De Deyn P, Mateo I, Franck A, Helisalmi S, Porcellini E, Hanon O, de Pancorbo MM, Lendon C, Dufouil C, Jaillard C, Leveillard T, Alvarez V, Bosco P, Mancuso M, Panza F, Nacmias B, Bossu P, Piccardi P, Annoni G, Seripa D, Galimberti D, Hannequin D, Licastro F, Soininen H, Ritchie K, Blanche H, Dartigues JF, Tzourio C, Gut I, Van Broeckhoven C, Alperovitch A, Lathrop M, Amouyel P. Genome-wide association study identifies variants at CLU and CR1 associated with Alzheimer’s disease. Nat Genet. 2009;41:1094–1099. doi: 10.1038/ng.439. [DOI] [PubMed] [Google Scholar]
  • 50.Harold D, Abraham R, Hollingworth P, Sims R, Gerrish A, Hamshere ML, Pahwa JS, Moskvina V, Dowzell K, Williams A, Jones N, Thomas C, Stretton A, Morgan AR, Lovestone S, Powell J, Proitsi P, Lupton MK, Brayne C, Rubinsztein DC, Gill M, Lawlor B, Lynch A, Morgan K, Brown KS, Passmore PA, Craig D, McGuinness B, Todd S, Holmes C, Mann D, Smith AD, Love S, Kehoe PG, Hardy J, Mead S, Fox N, Rossor M, Collinge J, Maier W, Jessen F, Schurmann B, van den Bussche H, Heuser I, Kornhuber J, Wiltfang J, Dichgans M, Frolich L, Hampel H, Hull M, Rujescu D, Goate AM, Kauwe JSK, Cruchaga C, Nowotny P, Morris JC, Mayo K, Sleegers K, Bettens K, Engelborghs S, De Deyn PP, Van Broeckhoven C, Livingston G, Bass NJ, Gurling H, McQuillin A, Gwilliam R, Deloukas P, Al-Chalabi A, Shaw CE, Tsolaki M, Singleton AB, Guerreiro R, Muhleisen TW, Nothen MM, Moebus S, Jockel KH, Klopp N, Wichmann HE, Carrasquillo MM, Pankratz VS, Younkin SG, Holmans PA, O’Donovan M, Owen MJ, Williams J. Genome-wide association study identifies variants at CLU and PICALM associated with Alzheimer’s disease. Nat Genet. 2009;41:1088–1093. doi: 10.1038/ng.440. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Hollingworth P, Harold D, Sims R, Gerrish A, Lambert JC, Carrasquillo MM, Abraham R, Hamshere ML, Pahwa JS, Moskvina V, Dowzell K, Jones N, Stretton A, Thomas C, Richards A, Ivanov D, Widdowson C, Chapman J, Lovestone S, Powell J, Proitsi P, Lupton MK, Brayne C, Rubinsztein DC, Gill M, Lawlor B, Lynch A, Brown KS, Passmore PA, Craig D, McGuinness B, Todd S, Holmes C, Mann D, Smith AD, Beaumont H, Warden D, Wilcock G, Love S, Kehoe PG, Hooper NM, Vardy ER, Hardy J, Mead S, Fox NC, Rossor M, Collinge J, Maier W, Jessen F, Ruther E, Schurmann B, Heun R, Kolsch H, van den Bussche H, Heuser I, Kornhuber J, Wiltfang J, Dichgans M, Frolich L, Hampel H, Gallacher J, Hull M, Rujescu D, Giegling I, Goate AM, Kauwe JSK, Cruchaga C, Nowotny P, Morris JC, Mayo K, Sleegers K, Bettens K, Engelborghs S, De Deyn PP, Van Broeckhoven C, Livingston G, Bass NJ, Gurling H, McQuillin A, Gwilliam R, Deloukas P, Al-Chalabi A, Shaw CE, Tsolaki M, Singleton AB, Guerreiro R, Muhleisen TW, Nothen MM, Moebus S, Jockel KH, Klopp N, Wichmann HE, Pankratz VS, Sando SB, Aasly JO, Barcikowska M, Wszolek ZK, Dickson DW, Graff-Radford NR, Petersen RC, van Duijn CM, Breteler MMB, Ikram MA, DeStefano AL, Fitzpatrick AL, Lopez O, Launer LJ, Seshadri S, Berr C, Campion D, Epelbaum J, Dartigues JF, Tzourio C, Alperovitch A, Lathrop M, Feulner TM, Friedrich P, Riehle C, Krawczak M, Schreiber S, Mayhaus M, Nicolhaus S, Wagenpfeil S, Steinberg S, Stefansson H, Stefansson K, Snaedal J, Bjornsson S, Jonsson PV, Chouraki V, Genier-Boley B, Hiltunen M, Soininen H, Combarros O, Zelenika D, Delepine M, Bullido MJ, Pasquier F, Mateo I, Frank-Garcia A, Porcellini E, Hanon O, Coto E, Alvarez V, Bosco P, Siciliano G, Mancuso M, Panza F, Solfrizzi V, Nacmias B, Sorbi S, Bossu P, Piccardi P, Arosio B, Annoni G, Seripa D, Pilotto A, Scarpini E, Galimberti D, Brice A, Hannequin D, Licastro F, Jones L, Holmans PA, Jonsson T, Riemenschneider M, Morgan K, Younkin SG, Owen MJ, O’Donovan M, Amouyel P, Williams J. Common variants at ABCA7, MS4A6A/MS4A4E, EPHA1, CD33 and CD2AP are associated with Alzheimer’s disease. Nat Genet. 2011 May;43:429–35. doi: 10.1038/ng.803. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Yu JT, Tan L. The Role of Clusterin in Alzheimer’s Disease: Pathways, Pathogenesis, and Therapy. Mol Neurobiol. 2012;45:314–326. doi: 10.1007/s12035-012-8237-1. [DOI] [PubMed] [Google Scholar]
  • 53.Youle RJ, Strasser A. The BCL-2 protein family: opposing activities that mediate cell death. Nat Rev Mol Cell Biol. 2008;9:47–59. doi: 10.1038/nrm2308. [DOI] [PubMed] [Google Scholar]
  • 54.Thambisetty M, An Y, Kinsey A, Koka D, Saleem M, Guntert A, Kraut M, Ferrucci L, Davatzikos C, Lovestone S, Resnick SM. Plasma clusterin concentration is associated with longitudinal brain atrophy in mild cognitive impairment. Neuroimage. 2012;59:212–217. doi: 10.1016/j.neuroimage.2011.07.056. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Sborgi L, Barrera-Vilarmau S, Obregon P, de Alba E. Characterization of a novel interaction between Bcl-2 members Diva and Harakiri. PLoS One. 2010;5:e15575. doi: 10.1371/journal.pone.0015575. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Bis JC, DeCarli C, Smith AV, van der Lijn F, Crivello F, Fornage M, Debette S, Shulman JM, Schmidt H, Srikanth V, Schuur M, Yu L, Choi SH, Sigurdsson S, Verhaaren BFJ, DeStefano AL, Lambert JC, Jack CR, Struchalin M, Stankovich J, Ibrahim-Verbaas CA, Fleischman D, Zijdenbos A, den Heijer T, Mazoyer B, Coker LH, Enzinger C, Danoy P, Amin N, Arfanakis K, van Buchem MA, de Bruijn RF, Beiser A, Dufouil C, Huang J, Cavalieri M, Thomson R, Niessen WJ, Chibnik LB, Gislason GK, Hofman A, Pikula A, Amouyel P, Freeman KB, Phan TG, Oostra BA, Stein JL, Medland SE, Vasquez AA, Hibar DP, Wright MJ, Franke B, Martin NG, Thompson PM, Nalls MA, Uitterlinden AG, Au R, Elbaz A, Beare RJ, van Swieten JC, Lopez OL, Harris TB, Chouraki V, Breteler MMB, De Jager PL, Becker JT, Vernooij MW, Knopman D, Fazekas F, Wolf PA, van der Lugt A, Gudnason V, Longstreth WT, Brown MA, Bennett DA, van Duijn CM, Mosley TH, Schmidt R, Tzourio C, Launer LJ, Ikram MA, Seshadri S. Common variants at 12q14 and 12q24 are associated with hippocampal volume. Nature Genetics. 2012;44:545–551. doi: 10.1038/ng.2237. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Jehle AW, Gardai SJ, Li S, Linsel-Nitschke P, Morimoto K, Janssen WJ, Vandivier RW, Wang N, Greenberg S, Dale BM, Qin C, Henson PM, Tall AR. ATP-binding cassette transporter A7 enhances phagocytosis of apoptotic cells and associated ERK signaling in macrophages. J Cell Biol. 2006;174:547–556. doi: 10.1083/jcb.200601030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Jenner P, Olanow CW. Understanding cell death in Parkinson’s disease. Ann Neurol. 1998;44:S72–84. doi: 10.1002/ana.410440712. [DOI] [PubMed] [Google Scholar]
  • 59.International Parkinson Disease Genomics Consortium. Nalls MA, Plagnol V, Hernandez DG, Sharma M, Sheerin UM, Saad M, Simon-Sanchez J, Schulte C, Lesage S, Sveinbjornsdottir S, Stefansson K, Martinez M, Hardy J, Heutink P, Brice A, Gasser T, Singleton AB, Wood NW. Imputation of sequence variants for identification of genetic risks for Parkinson’s disease: a meta-analysis of genome-wide association studies. Lancet. 2011;377:641–649. doi: 10.1016/S0140-6736(10)62345-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Xu J, Kao SY, Lee FJ, Song W, Jin LW, Yankner BA. Dopamine-dependent neurotoxicity of alpha-synuclein: a mechanism for selective neurodegeneration in Parkinson disease. Nat Med. 2002;8:600–606. doi: 10.1038/nm0602-600. [DOI] [PubMed] [Google Scholar]
  • 61.Iaccarino C, Crosio C, Vitale C, Sanna G, Carri MT, Barone P. Apoptotic mechanisms in mutant LRRK2-mediated cell death. Hum Mol Genet. 2007;16:1319–1326. doi: 10.1093/hmg/ddm080. [DOI] [PubMed] [Google Scholar]
  • 62.Tong Y, Yamaguchi H, Giaime E, Boyle S, Kopan R, Kelleher RJ 3rd, Shen J. Loss of leucine-rich repeat kinase 2 causes impairment of protein degradation pathways, accumulation of alpha-synuclein, and apoptotic cell death in aged mice. Proc Natl Acad Sci U S A. 2010;107:9879–9884. doi: 10.1073/pnas.1004676107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Qureshi HY, Paudel HK. Parkinsonian neurotoxin 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) and alpha-synuclein mutations promote Tau protein phosphorylation at Ser262 and destabilize microtubule cytoskeleton in vitro. J Biol Chem. 2011;286:5055–5068. doi: 10.1074/jbc.M110.178905. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Johnson BN, Berger AK, Cortese GP, Lavoie MJ. The ubiquitin E3 ligase parkin regulates the proapoptotic function of Bax. Proc Natl Acad Sci U S A. 2012;109:6283–6288. doi: 10.1073/pnas.1113248109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Sureda FX, Junyent F, Verdaguer E, Auladell C, Pelegri C, Vilaplana J, Folch J, Canudas AM, Zarate CB, Palles M, Camins A. Antiapoptotic drugs: a therapautic strategy for the prevention of neurodegenerative diseases. Curr Pharm Des. 2011;17:230–245. doi: 10.2174/138161211795049732. [DOI] [PubMed] [Google Scholar]
  • 66.Rohn TT. The role of caspases in Alzheimer’s disease; potential novel therapeutic opportunities. Apoptosis. 2010;15:1403–1409. doi: 10.1007/s10495-010-0463-2. [DOI] [PubMed] [Google Scholar]
  • 67.Tabares-Seisdedos R, Dumont N, Baudot A, Valderas JM, Climent J, Valencia A, Crespo-Facorro B, Vieta E, Gomez-Beneyto M, Martinez S, Rubenstein JL. No paradox, no progress: inverse cancer comorbidity in people with other complex diseases. Lancet Oncol. 2011;12:604–608. doi: 10.1016/S1470-2045(11)70041-9. [DOI] [PubMed] [Google Scholar]
  • 68.Holohan KN, Lahiri DK, Schneider BP, Foroud T, Saykin AJ. Functional microRNAs in Alzheimer’s disease and cancer: differential regulation of common mechanisms and pathway. Front Genet. 2012;3:323. doi: 10.3389/fgene.2012.00323. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Lipinski MM, Zheng B, Lu T, Yan Z, Py BF, Ng A, Xavier RJ, Li C, Yankner BA, Scherzer CR, Yuan J. Genome-wide analysis reveals mechanisms modulating autophagy in normal brain aging and in Alzheimer’s disease. Proc Natl Acad Sci U S A. 2010;107:14164–14169. doi: 10.1073/pnas.1009485107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Sidransky E, Lopez G. The link between the GBA gene and parkinsonism. Lancet Neurol. 2012;11:986–998. doi: 10.1016/S1474-4422(12)70190-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Cullen V, Sardi SP, Ng J, Xu YH, Sun Y, Tomlinson JJ, Kolodziej P, Kahn I, Saftig P, Woulfe J, Rochet JC, Glicksman MA, Cheng SH, Grabowski GA, Shihabuddin LS, Schlossmacher MG. Acid β-glucosidase mutants linked to gaucher disease, parkinson disease, and lewy body dementia alter α-synuclein processing. Ann Neurol. 2011;69:940–953. doi: 10.1002/ana.22400. [DOI] [PubMed] [Google Scholar]
  • 72.Yang DS, Kumar A, Stavrides P, Peterson J, Peterhoff CM, Pawlik M, Levy E, Cataldo AM, Nixon RA. Neuronal apoptosis and autophagy cross talk in aging PS/APP mice, a model of Alzheimer’s disease. Am J Pathol. 2008;173:665–681. doi: 10.2353/ajpath.2008.071176. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Kroemer G, Levine B. Autophagic cell death: the story of a misnomer. Nat Rev Mol Cell Biol. 2008;9:1004–1010. doi: 10.1038/nrm2527. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Harris H, Rubinsztein DC. Control of autophagy as a therapy for neurodegenerative disease. Nat Rev Neurol. 2012;8:108–117. doi: 10.1038/nrneurol.2011.200. [DOI] [PubMed] [Google Scholar]
  • 75.Atamna H, Frey WH 2nd. Mechanisms of mitochondrial dysfunction and energy deficiency in Alzheimer’s disease. Mitochondrion. 2007;7:297–310. doi: 10.1016/j.mito.2007.06.001. [DOI] [PubMed] [Google Scholar]
  • 76.De Strooper B, Scorrano L. Close encounter: mitochondria, endoplasmic reticulum and Alzheimer’s disease. EMBO J. 2012;31:4095–4097. doi: 10.1038/emboj.2012.279. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Chang S, ran Ma T, Miranda RD, Balestra ME, Mahley RW, Huang Y. Lipid- and receptor-binding regions of apolipoprotein E4 fragments act in concert to cause mitochondrial dysfunction and neurotoxicity. Proc Natl Acad Sci U S A. 2005;102:18694–18699. doi: 10.1073/pnas.0508254102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Bruck S, Huber TB, Ingham RJ, Kim K, Niederstrasser H, Allen PM, Pawson T, Cooper JA, Shaw AS. Identification of a novel inhibitory actin-capping protein binding motif in CD2-associated protein. J Biol Chem. 2006;281:19196–19203. doi: 10.1074/jbc.M600166200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.DuBoff B, Gotz J, Feany MB. Tau promotes neurodegeneration via DRP1 mislocalization in vivo. Neuron. 2012;75:618–632. doi: 10.1016/j.neuron.2012.06.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Iijima-Ando K, Sekiya M, Maruko-Otake A, Ohtake Y, Suzuki E, Lu B, Iijima KM. Loss of axonal mitochondria promotes tau-mediated neurodegeneration and Alzheimer’s disease-related tau phosphorylation via PAR-1. PLoS Genet. 2012;8:e1002918. doi: 10.1371/journal.pgen.1002918. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Krstic D, Knuesel I. Deciphering the mechanism underlying late-onset Alzheimer disease. Nat Rev Neurol. 2013;9:25–34. doi: 10.1038/nrneurol.2012.236. [DOI] [PubMed] [Google Scholar]
  • 82.Rupprecht R, Papadopoulos V, Rammes G, Baghai TC, Fan J, Akula N, Groyer G, Adams D, Schumacher M. Translocator protein (18 kDa) (TSPO) as a therapeutic target for neurological and psychiatric disorders. Nat Rev Drug Discov. 2010;9:971–988. doi: 10.1038/nrd3295. [DOI] [PubMed] [Google Scholar]
  • 83.Caselli RJ, Dueck AC, Huentelman MJ, Lutz MW, Saunders AM, Reiman EM, Roses AD. Longitudinal modeling of cognitive aging and the TOMM40 effect. Alzheimers Dement. 2012;8:490–495. doi: 10.1016/j.jalz.2011.11.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Henchcliffe C, Beal MF. Mitochondrial biology and oxidative stress in Parkinson disease pathogenesis. Nat Clin Pract Neurol. 2008;4:600–609. doi: 10.1038/ncpneuro0924. [DOI] [PubMed] [Google Scholar]
  • 85.Lesage S, Brice A. Role of mendelian genes in “sporadic” Parkinson’s disease. Parkinsonism Relat Disord. 2012;18(Suppl 1):S66–70. doi: 10.1016/S1353-8020(11)70022-0. [DOI] [PubMed] [Google Scholar]
  • 86.Palikaras K, Tavernarakis N. Mitophagy in neurodegeneration and aging. Front Genet. 2012;3:297. doi: 10.3389/fgene.2012.00297. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Liu S, Sawada T, Lee S, Yu W, Silverio G, Alapatt P, Millan I, Shen A, Saxton W, Kanao T, Takahashi R, Hattori N, Imai Y, Lu B. Parkinson’s disease-associated kinase PINK1 regulates Miro protein level and axonal transport of mitochondria. PLoS Genet. 2012;8:e1002537. doi: 10.1371/journal.pgen.1002537. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Thomas KJ, McCoy MK, Blackinton J, Beilina A, van der Brug M, Sandebring A, Miller D, Maric D, Cedazo-Minguez A, Cookson MR. DJ-1 acts in parallel to the PINK1/parkin pathway to control mitochondrial function and autophagy. Hum Mol Genet. 2011;20:40–50. doi: 10.1093/hmg/ddq430. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Langston JW, Ballard P, Tetrud JW, Irwin I. Chronic Parkinsonism in humans due to a product of meperidine-analog synthesis. Science. 1983;219:979–980. doi: 10.1126/science.6823561. [DOI] [PubMed] [Google Scholar]
  • 90.Blesa J, Phani S, Jackson-Lewis V, Przedborski S. Classic and new animal models of Parkinson’s disease. J Biomed Biotechnol. 2012;2012:845618. doi: 10.1155/2012/845618. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Jenner P. Parkinson’s disease, pesticides and mitochondrial dysfunction. Trends Neurosci. 2001;24:245–247. doi: 10.1016/s0166-2236(00)01789-6. [DOI] [PubMed] [Google Scholar]
  • 92.Yan MH, Wang X, Zhu X. Mitochondrial defects and oxidative stress in Alzheimer disease and Parkinson disease. Free Radic Biol Med. 2013 Sep;62:90–101. doi: 10.1016/j.freeradbiomed.2012.11.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Lin MT, Beal MF. Mitochondrial dysfunction and oxidative stress in neurodegenerative diseases. Nature. 2006;443:787–795. doi: 10.1038/nature05292. [DOI] [PubMed] [Google Scholar]
  • 94.Squier TC. Oxidative stress and protein aggregation during biological aging. Exp Gerontol. 2001;36:1539–1550. doi: 10.1016/s0531-5565(01)00139-5. [DOI] [PubMed] [Google Scholar]
  • 95.Kim RH, Smith PD, Aleyasin H, Hayley S, Mount MP, Pownall S, Wakeham A, You-Ten AJ, Kalia SK, Horne P, Westaway D, Lozano AM, Anisman H, Park DS, Mak TW. Hypersensitivity of DJ-1-deficient mice to 1-methyl-4-phenyl-1,2,3,6-tetrahydropyrindine (MPTP) and oxidative stress. Proc Natl Acad Sci U S A. 2005;102:5215–5220. doi: 10.1073/pnas.0501282102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Irrcher I, Aleyasin H, Seifert EL, Hewitt SJ, Chhabra S, Phillips M, Lutz AK, Rousseaux MWC, Bevilacqua L, Jahani-Asl A, Callaghan S, MacLaurin JG, Winklhofer KF, Rizzu P, Rippstein P, Kim RH, Chen CX, Fon EA, Slack RS, Harper ME, McBride HM, Mak TW, Park DS. Loss of the Parkinson’s disease-linked gene DJ-1 perturbs mitochondrial dynamics. Hum Mol Genet. 2010 Oct 1;19:3734–46. doi: 10.1093/hmg/ddq288. [DOI] [PubMed] [Google Scholar]
  • 97.Wu ZC, Yu JT, Li Y, Tan L. Clusterin in Alzheimer’s disease. Adv Clin Chem. 2012;56:155–173. doi: 10.1016/b978-0-12-394317-0.00011-x. [DOI] [PubMed] [Google Scholar]
  • 98.Allen M, Zou F, Chai HS, Younkin CS, Miles R, Nair AA, Crook JE, Pankratz VS, Carrasquillo MM, Rowley CN, Nguyen T, Ma L, Malphrus KG, Bisceglio G, Ortolaza AI, Palusak R, Middha S, Maharjan S, Georgescu C, Schultz D, Rakhshan F, Kolbert CP, Jen J, Sando SB, Aasly JO, Barcikowska M, Uitti RJ, Wszolek ZK, Ross OA, Petersen RC, Graff-Radford NR, Dickson DW, Younkin SG, Ertekin-Taner N. Glutathione S-transferase omega genes in Alzheimer and Parkinson disease risk, age-at-diagnosis and brain gene expression: an association study with mechanistic implications. Mol Neurodegener. 2012;7:13. doi: 10.1186/1750-1326-7-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.De Jager PL, Shulman JM, Chibnik LB, Keenan BT, Raj T, Wilson RS, Yu L, Leurgans SE, Tran D, Aubin C, Anderson CD, Biffi A, Corneveaux JJ, Huentelman MJ Alzheimer’s Disease Neuroimaging Initiative. Rosand J, Daly MJ, Myers AJ, Reiman EM, Bennett DA, Evans DA. A genome-wide scan for common variants affecting the rate of age-related cognitive decline. Neurobiol Aging. 2012;33:1017, e1011–1015. doi: 10.1016/j.neurobiolaging.2011.09.033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Bajaj A, Driver JA, Schernhammer ES. Parkinson’s disease and cancer risk: a systematic review and meta-analysis. Cancer Causes Control. 2010;21:697–707. doi: 10.1007/s10552-009-9497-6. [DOI] [PubMed] [Google Scholar]
  • 101.Driver JA, Beiser A, Au R, Kreger BE, Splansky GL, Kurth T, Kiel DP, Lu KP, Seshadri S, Wolf PA. Inverse association between cancer and Alzheimer’s disease: results from the Framingham Heart Study. BMJ. 2012;344:e1442. doi: 10.1136/bmj.e1442. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Devine MJ, Plun-Favreau H, Wood NW. Parkinson’s disease and cancer: two wars, one front. Nat Rev Cancer. 2011;11:812–823. doi: 10.1038/nrc3150. [DOI] [PubMed] [Google Scholar]
  • 103.Tollervey JR, Wang Z, Hortobagyi T, Witten JT, Zarnack K, Kayikci M, Clark TA, Schweitzer AC, Rot G, Curk T, Zupan B, Rogelj B, Shaw CE, Ule J. Analysis of alternative splicing associated with aging and neurodegeneration in the human brain. Genome Res. 2011;21:1572–1582. doi: 10.1101/gr.122226.111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Kraft C, Peter M, Hofmann K. Selective autophagy: ubiquitin-mediated recognition and beyond. Nat Cell Biol. 2010;12:836–841. doi: 10.1038/ncb0910-836. [DOI] [PubMed] [Google Scholar]
  • 105.Ferrari R, Hardy J, Momeni P. Frontotemporal dementia: from Mendelian genetics towards genome wide association studies. J Mol Neurosci. 2011;45:500–515. doi: 10.1007/s12031-011-9635-y. [DOI] [PubMed] [Google Scholar]
  • 106.Simpson JE, Ince PG, Shaw PJ, Heath PR, Raman R, Garwood CJ, Gelsthorpe C, Baxter L, Forster G, Matthews FE, Brayne C, Wharton SB MRC Cognitive Function and Ageing Neuropathology Study Group. Microarray analysis of the astrocyte transcriptome in the aging brain: relationship to Alzheimer’s pathology and APOE genotype. Neurobiol Aging. 2011;32:1795–1807. doi: 10.1016/j.neurobiolaging.2011.04.013. [DOI] [PubMed] [Google Scholar]
  • 107.Kopito RR, Ron D. Conformational disease. Nat Cell Biol. 2000;2:E207–209. doi: 10.1038/35041139. [DOI] [PubMed] [Google Scholar]
  • 108.Scorrano L, Oakes SA, Opferman JT, Cheng EH, Sorcinelli MD, Pozzan T, Korsmeyer SJ. BAX and BAK regulation of endoplasmic reticulum Ca2+: a control point for apoptosis. Science. 2003;300:135–139. doi: 10.1126/science.1081208. [DOI] [PubMed] [Google Scholar]
  • 109.Smith DL, Pozueta J, Gong B, Arancio O, Shelanski M. Reversal of long-term dendritic spine alterations in Alzheimer disease models. Proc Natl Acad Sci U S A. 2009;106:16877–16882. doi: 10.1073/pnas.0908706106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Gong B, Cao Z, Zheng P, Vitolo OV, Liu S, Staniszewski A, Moolman D, Zhang H, Shelanski M, Arancio O. Ubiquitin hydrolase Uch-L1 rescues beta-amyloid-induced decreases in synaptic function and contextual memory. Cell. 2006;126:775–788. doi: 10.1016/j.cell.2006.06.046. [DOI] [PubMed] [Google Scholar]
  • 111.Khandelwal PJ, Herman AM, Hoe HS, Rebeck GW, Moussa CE. Parkin mediates beclin-dependent autophagic clearance of defective mitochondria and ubiquitinated Abeta in AD models. Hum Mol Genet. 2011;20:2091–2102. doi: 10.1093/hmg/ddr091. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Huynh DP, Scoles DR, Nguyen D, Pulst SM. The autosomal recessive juvenile Parkinson disease gene product, parkin, interacts with and ubiquitinates synaptotagmin XI. Hum Mol Genet. 2003;12:2587–2597. doi: 10.1093/hmg/ddg269. [DOI] [PubMed] [Google Scholar]
  • 113.Rana A, Rera M, Walker DW. Parkin overexpression during aging reduces proteotoxicity, alters mitochondrial dynamics, and extends lifespan. Proc Natl Acad Sci U S A. 2013;110:8638–8643. doi: 10.1073/pnas.1216197110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Horwitz AR. The origins of the molecular era of adhesion research. Nat Rev Mol Cell Biol. 2012;13:805–811. doi: 10.1038/nrm3473. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Blalock EM, Geddes JW, Chen KC, Porter NM, Markesbery WR, Landfield PW. Incipient Alzheimer’s disease: microarray correlation analyses reveal major transcriptional and tumor suppressor responses. Proc Natl Acad Sci U S A. 2004;101:2173–2178. doi: 10.1073/pnas.0308512100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Grunblatt E, Mandel S, Jacob-Hirsch J, Zeligson S, Amariglo N, Rechavi G, Li J, Ravid R, Roggendorf W, Riederer P, Youdim MB. Gene expression profiling of parkinsonian substantia nigra pars compacta; alterations in ubiquitin-proteasome, heat shock protein, iron and oxidative stress regulated proteins, cell adhesion/cellular matrix and vesicle trafficking genes. J Neural Transm. 2004;111:1543–1573. doi: 10.1007/s00702-004-0212-1. [DOI] [PubMed] [Google Scholar]
  • 117.Miller RM, Kiser GL, Kaysser-Kranich TM, Lockner RJ, Palaniappan C, Federoff HJ. Robust dysregulation of gene expression in substantia nigra and striatum in Parkinson’s disease. Neurobiol Dis. 2006;21:305–313. doi: 10.1016/j.nbd.2005.07.010. [DOI] [PubMed] [Google Scholar]
  • 118.Xu PT, Li YJ, Qin XJ, Scherzer CR, Xu H, Schmechel DE, Hulette CM, Ervin J, Gullans SR, Haines J, Pericak-Vance MA, Gilbert JR. Differences in apolipoprotein E3/3 and E4/4 allele-specific gene expression in hippocampus in Alzheimer disease. Neurobiol Dis. 2006;21:256–275. doi: 10.1016/j.nbd.2005.07.004. [DOI] [PubMed] [Google Scholar]
  • 119.Durakoglugil MS, Chen Y, White CL, Kavalali ET, Herz J. Reelin signaling antagonizes beta-amyloid at the synapse. Proc Natl Acad Sci U S A. 2009;106:15938–15943. doi: 10.1073/pnas.0908176106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Chen Y, Durakoglugil MS, Xian X, Herz J. ApoE4 reduces glutamate receptor function and synaptic plasticity by selectively impairing ApoE receptor recycling. Proc Natl Acad Sci U S A. 2010;107:12011–12016. doi: 10.1073/pnas.0914984107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Herring A, Donath A, Steiner KM, Widera MP, Hamzehian S, Kanakis D, Kolble K, ElAli A, Hermann DM, Paulus W, Keyvani K. Reelin depletion is an early phenomenon of Alzheimer’s pathology. J Alzheimers Dis. 2012;30:963–979. doi: 10.3233/JAD-2012-112069. [DOI] [PubMed] [Google Scholar]
  • 122.Kramer PL, Xu H, Woltjer RL, Westaway SK, Clark D, Erten-Lyons D, Kaye JA, Welsh-Bohmer KA, Troncoso JC, Markesbery WR, Petersen RC, Turner RS, Kukull WA, Bennett DA, Galasko D, Morris JC, Ott J. Alzheimer disease pathology in cognitively healthy elderly: a genome-wide study. Neurobiol Aging. 2011;32:2113–2122. doi: 10.1016/j.neurobiolaging.2010.01.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Kopan R, Ilagan MX. Gamma-secretase: proteasome of the membrane? Nat Rev Mol Cell Biol. 2004;5:499–504. doi: 10.1038/nrm1406. [DOI] [PubMed] [Google Scholar]
  • 124.Chen KP, Dou F. Selective interaction of amyloid precursor protein with different isoforms of neural cell adhesion molecule. J Mol Neurosci. 2012;46:203–209. doi: 10.1007/s12031-011-9578-3. [DOI] [PubMed] [Google Scholar]
  • 125.Wang Y, West JD, MaGee TR, Mcdonald BC, Risacher SL, Shen L, Ramanan VK, Kim S, O’Neill DP, Farlow MR, Ghetti B, Saykin AJ. Hippocampal subfield atrophy on 3T MRI in prodromal Alzheimer’s disease and older adults with cognitive complaints. . 2012 Poster Presentation, Alzheimer’s Association International Conference (AAIC) in Vancouver, British Colombia, Canada, July 14-19, 2012. [Google Scholar]
  • 126.Saykin AJ, Shen L, Foroud TM, Potkin SG, Swaminathan S, Kim S, Risacher SL, Nho K, Huentelman MJ, Craig DW, Thompson PM, Stein JL, Moore JH, Farrer LA, Green RC, Bertram L, Jack CR Jr, Weiner MW Alzheimer’s Disease Neuroimaging Initiative. Alzheimer’s Disease Neuroimaging Initiative biomarkers as quantitative phenotypes: Genetics core aims, progress, and plans. Alzheimers Dement. 2010;6:265–273. doi: 10.1016/j.jalz.2010.03.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Edwards YJ, Beecham GW, Scott WK, Khuri S, Bademci G, Tekin D, Martin ER, Jiang Z, Mash DC, ffrench-Mullen J, Pericak-Vance MA, Tsinoremas N, Vance JM. Identifying consensus disease pathways in Parkinson’s disease using an integrative systems biology approach. PLoS One. 2011;6:e16917. doi: 10.1371/journal.pone.0016917. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Liu G, Jiang Y, Wang P, Feng R, Jiang N, Chen X, Song H, Chen Z. Cell adhesion molecules contribute to Alzheimer’s disease: multiple pathway analyses of two genome-wide association studies. J Neurochem. 2012;120:190–198. doi: 10.1111/j.1471-4159.2011.07547.x. [DOI] [PubMed] [Google Scholar]
  • 129.Holmans P. Statistical methods for pathway analysis of genome-wide data for association with complex genetic traits. Adv Genet. 2010;72:141–179. doi: 10.1016/B978-0-12-380862-2.00007-2. [DOI] [PubMed] [Google Scholar]
  • 130.Zhang M. Endocytic mechanisms and drug discovery in neurodegenerative diseases. Front Biosci. 2008;13:6086–6105. doi: 10.2741/3140. [DOI] [PubMed] [Google Scholar]
  • 131.Jiang Q, Lee CY, Mandrekar S, Wilkinson B, Cramer P, Zelcer N, Mann K, Lamb B, Willson TM, Collins JL, Richardson JC, Smith JD, Comery TA, Riddell D, Holtzman DM, Tontonoz P, Landreth GE. ApoE promotes the proteolytic degradation of Abeta. Neuron. 2008;58:681–693. doi: 10.1016/j.neuron.2008.04.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Pottier C, Hannequin D, Coutant S, Rovelet-Lecrux A, Wallon D, Rousseau S, Legallic S, Paquet C, Bombois S, Pariente J, Thomas-Anterion C, Michon A, Croisile B, Etcharry-Bouyx F, Berr C, Dartigues JF, Amouyel P, Dauchel H, Boutoleau-Bretonniere C, Thauvin C, Frebourg T, Lambert JC, Campion D, Collaborators PG. High frequency of potentially pathogenic SORL1 mutations in autosomal dominant early- onset Alzheimer disease. Mol Psychiatry. 2012;17:875–879. doi: 10.1038/mp.2012.15. [DOI] [PubMed] [Google Scholar]
  • 133.Reitz C, Cheng R, Rogaeva E, Lee JH, Tokuhiro S, Zou F, Bettens K, Sleegers K, Tan EK, Kimura R, Shibata N, Arai H, Kamboh MI, Prince JA, Maier W, Riemenschneider M, Owen M, Harold D, Hollingworth P, Cellini E, Sorbi S, Nacmias B, Takeda M, Pericak-Vance MA, Haines JL, Younkin S, Williams J, van Broeckhoven C, Farrer LA, St George-Hyslop PH, Mayeux R Genetic and Environmental Risk in Alzheimer Disease Consortium. Meta-analysis of the association between variants in SORL1 and Alzheimer disease. Arch Neurol. 2011;68:99–106. doi: 10.1001/archneurol.2010.346. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Rogaeva E, Meng Y, Lee JH, Gu Y, Kawarai T, Zou F, Katayama T, Baldwin CT, Cheng R, Hasegawa H, Chen F, Shibata N, Lunetta KL, Pardossi-Piquard R, Bohm C, Wakutani Y, Cupples LA, Cuenco KT, Green RC, Pinessi L, Rainero I, Sorbi S, Bruni A, Duara R, Friedland RP, Inzelberg R, Hampe W, Bujo H, Song YQ, Andersen OM, Willnow TE, Graff-Radford N, Petersen RC, Dickson D, Der SD, Fraser PE, Schmitt-Ulms G, Younkin S, Mayeux R, Farrer LA, St George-Hyslop P. The neuronal sortilin-related receptor SORL1 is genetically associated with Alzheimer disease. Nat Genet. 2007;39:168–177. doi: 10.1038/ng1943. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Golde TE, Estus S, Younkin LH, Selkoe DJ, Younkin SG. Processing of the amyloid protein precursor to potentially amyloidogenic derivatives. Science. 1992;255:728–730. doi: 10.1126/science.1738847. [DOI] [PubMed] [Google Scholar]
  • 136.Alegre-Abarrategui J, Wade-Martins R. Parkinson disease, LRRK2 and the endocytic-autophagic pathway. Autophagy. 2009;5:1208–1210. doi: 10.4161/auto.5.8.9894. [DOI] [PubMed] [Google Scholar]
  • 137.Alegre-Abarrategui J, Christian H, Lufino MM, Mutihac R, Venda LL, Ansorge O, Wade-Martins R. LRRK2 regulates autophagic activity and localizes to specific membrane microdomains in a novel human genomic reporter cellular model. Hum Mol Genet. 2009;18:4022–4034. doi: 10.1093/hmg/ddp346. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Matta S, Van Kolen K, da Cunha R, van den Bogaart G, Mandemakers W, Miskiewicz K, De Bock PJ, Morais VA, Vilain S, Haddad D, Delbroek L, Swerts J, Chavez-Gutierrez L, Esposito G, Daneels G, Karran E, Holt M, Gevaert K, Moechars DW, De Strooper B, Verstreken P. LRRK2 controls an EndoA phosphorylation cycle in synaptic endocytosis. Neuron. 2012;75:1008–1021. doi: 10.1016/j.neuron.2012.08.022. [DOI] [PubMed] [Google Scholar]
  • 139.Cramer PE, Cirrito JR, Wesson DW, Lee CYD, Karlo JC, Zinn AE, Casali BT, Restivo JL, Goebel WD, James MJ, Brunden KR, Wilson DA, Landreth GE. ApoE-Directed Therapeutics Rapidly Clear β-Amyloid and Reverse Deficits in AD Mouse Models. Science. 2012 Mar 23;335:1503–6. doi: 10.1126/science.1217697. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Treusch S, Hamamichi S, Goodman JL, Matlack KES, Chung CY, Baru V, Shulman JM, Parrado A, Bevis BJ, Valastyan JS, Han H, Lindhagen-Persson M, Reiman EM, Evans DA, Bennett DA, Olofsson A, DeJager PL, Tanzi RE, Caldwell KA, Caldwell GA, Lindquist S. Functional Links Between Aβ Toxicity, Endocytic Trafficking, and Alzheimer’s Disease Risk Factors in Yeast. Science. 2011 Dec 2;334:1241–5. doi: 10.1126/science.1213210. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Lee DW, Zhao X, Zhang F, Eisenberg E, Greene LE. Depletion of GAK/auxilin 2 inhibits receptor-mediated endocytosis and recruitment of both clathrin and clathrin adaptors. J Cell Sci. 2005;118:4311–4321. doi: 10.1242/jcs.02548. [DOI] [PubMed] [Google Scholar]
  • 142.Dumitriu A, Pacheco CD, Wilk JB, Strathearn KE, Latourelle JC, Goldwurm S, Pezzoli G, Rochet JC, Lindquist S, Myers RH. Cyclin-G-associated kinase modifies alpha-synuclein expression levels and toxicity in Parkinson’s disease: results from the GenePD Study. Hum Mol Genet. 2011;20:1478–1487. doi: 10.1093/hmg/ddr026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Henderson DM, Conner SD. A novel AAK1 splice variant functions at multiple steps of the endocytic pathway. Mol Biol Cell. 2007;18:2698–2706. doi: 10.1091/mbc.E06-09-0831. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Latourelle JC, Pankratz N, Dumitriu A, Wilk JB, Goldwurm S, Pezzoli G, Mariani CB, DeStefano AL, Halter C, Gusella JF, Nichols WC, Myers RH, Foroud T PROGENI Investigators, Coordinators and Molecular Genetic Laboratories; GenePD Investigators, Coordinators and Molecular Genetic Laboratories. Genomewide association study for onset age in Parkinson disease. BMC Med Genet. 2009;10:98. doi: 10.1186/1471-2350-10-98. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Davies P, Maloney AJ. Selective loss of central cholinergic neurons in Alzheimer’s disease. Lancet. 1976;2:1403. doi: 10.1016/s0140-6736(76)91936-x. [DOI] [PubMed] [Google Scholar]
  • 146.Hoehn MM, Yahr MD. Parkinsonism: onset, progression and mortality. Neurology. 1967;17:427–442. doi: 10.1212/wnl.17.5.427. [DOI] [PubMed] [Google Scholar]
  • 147.Edwards RH. Neural degeneration and the transport of neurotransmitters. Ann Neurol. 1993;34:638–645. doi: 10.1002/ana.410340504. [DOI] [PubMed] [Google Scholar]
  • 148.Corbett A, Pickett J, Burns A, Corcoran J, Dunnett SB, Edison P, Hagan JJ, Holmes C, Jones E, Katona C, Kearns I, Kehoe P, Mudher A, Passmore A, Shepherd N, Walsh F, Ballard C. Drug repositioning for Alzheimer’s disease. Nat Rev Drug Discov. 2012;11:833–846. doi: 10.1038/nrd3869. [DOI] [PubMed] [Google Scholar]
  • 149.Antonini A, Albin RL. Dopaminergic treatment and nonmotor features of Parkinson disease:The horse lives. Neurology. 2013;80:784–785. doi: 10.1212/WNL.0b013e318285c16d. [DOI] [PubMed] [Google Scholar]
  • 150.Ho A, Shen J. Presenilins in synaptic function and disease. Trends Mol Med. 2011;17:617–624. doi: 10.1016/j.molmed.2011.06.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Sweatt JD. Mechanisms of Memory. 2nd Edition. Academic Press; 2009. [Google Scholar]
  • 152.Wu K, O’Keeffe D, Politis M, O’Keeffe GC, Robbins TW, Bose SK, Brooks DJ, Piccini P, Barker RA. The catechol-O-methyltransferase Val(158)Met polymorphism modulates fronto-cortical dopamine turnover in early Parkinson’s disease: a PET study. Brain. 2012;135:2449–2457. doi: 10.1093/brain/aws157. [DOI] [PubMed] [Google Scholar]
  • 153.Serretti A, Olgiati P. Catechol-o-methyltransferase and Alzheimer’s disease: a review of biological and genetic findings. CNS Neurol Disord Drug Targets. 2012;11:299–305. doi: 10.2174/187152712800672472. [DOI] [PubMed] [Google Scholar]
  • 154.Spence RD, Voskuhl RR. Neuroprotective effects of estrogens and androgens in CNS inflammation and neurodegeneration. Front Neuroendocrinol. 2012;33:105–115. doi: 10.1016/j.yfrne.2011.12.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Goldstein M, Lieberman A. The role of the regulatory enzymes of catecholamine synthesis in Parkinson’s disease. Neurology. 1992;42:8–12. discussion 41-18. [PubMed] [Google Scholar]
  • 156.Gainetdinov RR, Fumagalli F, Jones SR, Caron MG. Dopamine transporter is required for in vivo MPTP neurotoxicity: evidence from mice lacking the transporter. J Neurochem. 1997;69:1322–1325. doi: 10.1046/j.1471-4159.1997.69031322.x. [DOI] [PubMed] [Google Scholar]
  • 157.Baranzini SE, Srinivasan R, Khankhanian P, Okuda DT, Nelson SJ, Matthews PM, Hauser SL, Oksenberg JR, Pelletier D. Genetic variation influences glutamate concentrations in brains of patients with multiple sclerosis. Brain. 2010;133:2603–2611. doi: 10.1093/brain/awq192. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Mayeux R. Clinical practice. Early Alzheimer’s disease. N Engl J Med. 2010;362:2194–2201. doi: 10.1056/NEJMcp0910236. [DOI] [PubMed] [Google Scholar]
  • 159.Alkalay A, Rabinovici GD, Zimmerman G, Agarwal N, Kaufer D, Miller BL, Jagust WJ, Soreq H. Plasma acetylcholinesterase activity correlates with intracerebral beta-amyloid load. Curr Alzheimer Res. 2013;10:48–56. [PMC free article] [PubMed] [Google Scholar]
  • 160.Ramanan VK, Risacher SL, Nho K, Kim S, Swaminathan S, Shen L, Foroud TM, Hakonarson H, Huentelman MJ, Aisen PS, Petersen RC, Green RC, Jack CR, Koeppe RA, Jagust WJ, Weiner MW, Saykin AJ. APOE and BCHE as modulators of cerebral amyloid deposition: a florbetapir PET genome-wide association study. Mol Psychiatry. 2013 Feb 19; doi: 10.1038/mp.2013.19. [Epub ahead of print] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Mesulam MM, Geula C. Butyrylcholinesterase reactivity differentiates the amyloid plaques of aging from those of dementia. Ann Neurol. 1994;36:722–727. doi: 10.1002/ana.410360506. [DOI] [PubMed] [Google Scholar]
  • 162.Darvesh S, Hopkins DA, Geula C. Neurobiology of butyrylcholinesterase. Nat Rev Neurosci. 2003;4:131–138. doi: 10.1038/nrn1035. [DOI] [PubMed] [Google Scholar]
  • 163.Darreh-Shori T, Forsberg A, Modiri N, Andreasen N, Blennow K, Kamil C, Ahmed H, Almkvist O, Långström B, Nordberg A. Differential levels of apolipoprotein E and butyrylcho-linesterase show strong association with pathological signs of Alzheimer’s disease in the brain in vivo. Neurobiol Aging. 2011 Dec;32:2320, e15–32. doi: 10.1016/j.neurobiolaging.2010.04.028. [DOI] [PubMed] [Google Scholar]
  • 164.Darvesh S, Cash MK, Reid GA, Martin E, Mitnitski A, Geula C. Butyrylcholinesterase is associated with beta-amyloid plaques in the transgenic APPSWE/PSEN1dE9 mouse model of Alzheimer disease. J Neuropathol Exp Neurol. 2012;71:2–14. doi: 10.1097/NEN.0b013e31823cc7a6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165.Lane RM, He Y. Butyrylcholinesterase genotype and gender influence Alzheimer’s disease phenotype. Alzheimers Dement. 2013 Mar;9:e1–73. doi: 10.1016/j.jalz.2010.12.005. [DOI] [PubMed] [Google Scholar]
  • 166.Sabbagh MN, Farlow MR, Relkin N, Beach TG. Do cholinergic therapies have disease-modifying effects in Alzheimer’s disease? Alzheimers Dement. 2006 Apr;2:118–25. doi: 10.1016/j.jalz.2006.02.001. [DOI] [PubMed] [Google Scholar]
  • 167.Farlow MR, Miller ML, Pejovic V. Treatment options in Alzheimer’s disease: maximizing benefit, managing expectations. Dement Geriatr Cogn Disord. 2008;25:408–422. doi: 10.1159/000122962. [DOI] [PubMed] [Google Scholar]
  • 168.Caughey B, Baron GS. Prions and their partners in crime. Nature. 2006;443:803–810. doi: 10.1038/nature05294. [DOI] [PubMed] [Google Scholar]
  • 169.Prusiner SB. Scrapie prions. Annu Rev Microbiol. 1989;43:345–374. doi: 10.1146/annurev.mi.43.100189.002021. [DOI] [PubMed] [Google Scholar]
  • 170.Prusiner SB. Cell biology. A unifying role for prions in neurodegenerative diseases. Science. 2012;336:1511–1513. doi: 10.1126/science.1222951. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Dermaut B, Croes EA, Rademakers R, Van den Broeck M, Cruts M, Hofman A, van Duijn CM, Van Broeckhoven C. PRNP Val129 homozygosity increases risk for early-onset Alzheimer’s disease. Ann Neurol. 2003;53:409–412. doi: 10.1002/ana.10507. [DOI] [PubMed] [Google Scholar]
  • 172.Riemenschneider M, Klopp N, Xiang W, Wagenpfeil S, Vollmert C, Muller U, Forstl H, Illig T, Kretzschmar H, Kurz A. Prion protein codon 129 polymorphism and risk of Alzheimer disease. Neurology. 2004;63:364–366. doi: 10.1212/01.wnl.0000130198.72589.69. [DOI] [PubMed] [Google Scholar]
  • 173.Gossrau G, Herting B, Mockel S, Kempe A, Koch R, Reichmann H, Lampe JB. Analysis of the polymorphic prion protein gene codon 129 in idiopathic Parkinson’s disease. J Neural Transm. 2006;113:331–337. doi: 10.1007/s00702-005-0329-x. [DOI] [PubMed] [Google Scholar]
  • 174.Combarros O, Sanchez-Guerra M, Llorca J, Alvarez-Arcaya A, Berciano J, Pena N, Fernandez-Viadero C. Polymorphism at codon 129 of the prion protein gene is not associated with sporadic AD. Neurology. 2000;55:593–595. doi: 10.1212/wnl.55.4.593. [DOI] [PubMed] [Google Scholar]
  • 175.Jeong BH, Lee KH, Lee YJ, Kim Y, Choi EK, Kim YH, Cho YS, Carp R, Kim YS. Lack of association between PRNP 1368 polymorphism and Alzheimer’s disease or vascular dementia. BMC Med Genet. 2009;10:32. doi: 10.1186/1471-2350-10-32. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176.Mead S, Poulter M, Uphill J, Beck J, Whitfield J, Webb TEF, Campbell T, Adamson G, Deriziotis P, Tabrizi SJ, Hummerich H, Verzilli C, Alpers MP, Whittaker JC, Collinge J. Genetic risk factors for variant Creutzfeldt–Jakob disease: a genome-wide association study. Lancet Neurol. 2009;8:57–66. doi: 10.1016/S1474-4422(08)70265-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Mead S, Uphill J, Beck J, Poulter M, Campbell T, Lowe J, Adamson G, Hummerich H, Klopp N, Rückert IM, Wichmann HE, Azazi D, Plagnol V, Pako WH, Whitfield J, Alpers MP, Whittaker J, Balding DJ, Zerr I, Kretzschmar H, Collinge J. Genome-wide association study in multiple human prion diseases suggests genetic risk factors additional to PRNP. Hum Mol Genet. 2012 Apr 15;21:1897–906. doi: 10.1093/hmg/ddr607. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 178.Lukic A, Mead S. Genome wide association studies and prion disease. Prion. 2011;5:154–160. doi: 10.4161/pri.5.3.16892. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Lansbury PT, Lashuel HA. A century-old debate on protein aggregation and neurodegeneration enters the clinic. Nature. 2006;443:774–779. doi: 10.1038/nature05290. [DOI] [PubMed] [Google Scholar]
  • 180.Swardfager W, Lanctot K, Rothenburg L, Wong A, Cappell J, Herrmann N. A meta-analysis of cytokines in Alzheimer’s disease. Biol Psychiatry. 2010;68:930–941. doi: 10.1016/j.biopsych.2010.06.012. [DOI] [PubMed] [Google Scholar]
  • 181.Akama KT, Van Eldik LJ. Beta-amyloid stimulation of inducible nitric-oxide synthase in astrocytes is interleukin-1beta- and tumor necrosis factor-alpha (TNFalpha)-dependent, and involves a TNFalpha receptor-associated factor- and NFkappaB-inducing kinase-dependent signaling mechanism. J Biol Chem. 2000;275:7918–7924. doi: 10.1074/jbc.275.11.7918. [DOI] [PubMed] [Google Scholar]
  • 182.Blum-Degen D, Muller T, Kuhn W, Gerlach M, Przuntek H, Riederer P. Interleukin-1 beta and interleukin-6 are elevated in the cerebrospinal fluid of Alzheimer’s and de novo Parkinson’s disease patients. Neurosci Lett. 1995;202:17–20. doi: 10.1016/0304-3940(95)12192-7. [DOI] [PubMed] [Google Scholar]
  • 183.Hirsch EC, Hunot S. Neuroinflammation in Parkinson’s disease: a target for neuroprotection? Lancet Neurol. 2009;8:382–397. doi: 10.1016/S1474-4422(09)70062-6. [DOI] [PubMed] [Google Scholar]
  • 184.Stewart WF, Kawas C, Corrada M, Metter EJ. Risk of Alzheimer’s disease and duration of NSAID use. Neurology. 1997;48:626–632. doi: 10.1212/wnl.48.3.626. [DOI] [PubMed] [Google Scholar]
  • 185.Wahner AD, Bronstein JM, Bordelon YM, Ritz B. Nonsteroidal anti-inflammatory drugs may protect against Parkinson disease. Neurology. 2007;69:1836–1842. doi: 10.1212/01.wnl.0000279519.99344.ad. [DOI] [PubMed] [Google Scholar]
  • 186.Prinz M, Priller J, Sisodia SS, Ransohoff RM. Heterogeneity of CNS myeloid cells and their roles in neurodegeneration. Nat Neurosci. 2011;14:1227–1235. doi: 10.1038/nn.2923. [DOI] [PubMed] [Google Scholar]
  • 187.Lucin KM, Wyss-Coray T. Immune activation in brain aging and neurodegeneration: too much or too little? Neuron. 2009;64:110–122. doi: 10.1016/j.neuron.2009.08.039. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Lee EJ, Woo MS, Moon PG, Baek MC, Choi IY, Kim WK, Junn E, Kim HS. α-Synuclein Activates Microglia by Inducing the Expressions of Matrix Metalloproteinases and the Subsequent Activation of Protease-Activated Receptor-1. J Immunol. 2010 Jul 1;185:615–23. doi: 10.4049/jimmunol.0903480. [DOI] [PubMed] [Google Scholar]
  • 189.Halle A, Hornung V, Petzold GC, Stewart CR, Monks BG, Reinheckel T, Fitzgerald KA, Latz E, Moore KJ, Golenbock DT. The NALP3 inflammasome is involved in the innate immune response to amyloid-[beta] . Nat Immunol. 2008;9:857–865. doi: 10.1038/ni.1636. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190.Kitazawa M, Cheng D, Tsukamoto MR, Koike MA, Wes PD, Vasilevko V, Cribbs DH, LaFerla FM. Blocking IL-1 Signaling Rescues Cognition, Attenuates Tau Pathology, and Restores Neuronal β-Catenin Pathway Function in an Alzheimer’s Disease Model. J Immunol. 2011 Dec 15;187:6539–49. doi: 10.4049/jimmunol.1100620. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Pott Godoy MC, Tarelli R, Ferrari CC, Sarchi MI, Pitossi FJ. Central and systemic IL-1 exacerbates neurodegeneration and motor symptoms in a model of Parkinson’s disease. Brain. 2008;131:1880–1894. doi: 10.1093/brain/awn101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.Heneka MT, Kummer MP, Stutz A, Delekate A, Schwartz S, Vieira-Saecker A, Griep A, Axt D, Remus A, Tzeng TC, Gelpi E, Halle A, Korte M, Latz E, Golenbock DT. NLRP3 is activated in Alzheimer’s disease and contributes to pathology in APP/PS1 mice. Nature. 2013;493:674–678. doi: 10.1038/nature11729. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 193.White GE, Greaves DR. Fractalkine: a survivor’s guide: chemokines as antiapoptotic mediators. Arterioscler Thromb Vasc Biol. 2012;32:589–594. doi: 10.1161/ATVBAHA.111.237412. [DOI] [PubMed] [Google Scholar]
  • 194.Cardona AE, Pioro EP, Sasse ME, Kostenko V, Cardona SM, Dijkstra IM, Huang D, Kidd G, Dombrowski S, Dutta R, Lee JC, Cook DN, Jung S, Lira SA, Littman DR, Ransohoff RM. Control of microglial neurotoxicity by the fractalkine receptor. Nat Neurosci. 2006;9:917–924. doi: 10.1038/nn1715. [DOI] [PubMed] [Google Scholar]
  • 195.Bamberger ME, Harris ME, McDonald DR, Husemann J, Landreth GE. A cell surface receptor complex for fibrillar beta-amyloid mediates microglial activation. J Neurosci. 2003;23:2665–2674. doi: 10.1523/JNEUROSCI.23-07-02665.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 196.El Khoury J, Toft M, Hickman SE, Means TK, Terada K, Geula C, Luster AD. Ccr2 deficiency impairs microglial accumulation and accelerates progression of Alzheimer-like disease. Nat Med. 2007;13:432–438. doi: 10.1038/nm1555. [DOI] [PubMed] [Google Scholar]
  • 197.Heneka MT, Nadrigny F, Regen T, Martinez-Hernandez A, Dumitrescu-Ozimek L, Terwel D, Jardanhazi-Kurutz D, Walter J, Kirchhoff F, Hanisch UK, Kummer MP. Locus ceruleus controls Alzheimer’s disease pathology by modulating microglial functions through norepinephrine. Proc Natl Acad Sci U S A. 2010;107:6058–6063. doi: 10.1073/pnas.0909586107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 198.Ramos EM, Lin MT, Larson EB, Maezawa I, Tseng LH, Edwards KL, Schellenberg GD, Hansen JA, Kukull WA, Jin LW. Tumor necrosis factor alpha and interleukin 10 promoter region polymorphisms and risk of late-onset Alzheimer disease. Arch Neurol. 2006;63:1165–1169. doi: 10.1001/archneur.63.8.1165. [DOI] [PubMed] [Google Scholar]
  • 199.Wahner AD, Sinsheimer JS, Bronstein JM, Ritz B. Inflammatory cytokine gene polymorphisms and increased risk of Parkinson disease. Arch Neurol. 2007;64:836–840. doi: 10.1001/archneur.64.6.836. [DOI] [PubMed] [Google Scholar]
  • 200.Guerreiro R, Wojtas A, Bras J, Carrasquillo M, Rogaeva E, Majounie E, Cruchaga C, Sassi C, Kauwe JS, Younkin S, Hazrati L, Collinge J, Pocock J, Lashley T, Williams J, Lambert JC, Amouyel P, Goate A, Rademakers R, Morgan K, Powell J, St George-Hyslop P, Singleton A, Hardy J Alzheimer Genetic Analysis Group. TREM2 variants in Alzheimer’s disease. N Engl J Med. 2013;368:117–127. doi: 10.1056/NEJMoa1211851. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201.Jonsson T, Stefansson H, Steinberg S, Jonsdottir I, Jonsson PV, Snaedal J, Bjornsson S, Huttenlocher J, Levey AI, Lah JJ, Rujescu D, Hampel H, Giegling I, Andreassen OA, Engedal K, Ulstein I, Djurovic S, Ibrahim-Verbaas C, Hofman A, Ikram MA, van Duijn CM, Thorsteinsdottir U, Kong A, Stefansson K. Variant of TREM2 associated with the risk of Alzheimer’s disease. N Engl J Med. 2013;368:107–116. doi: 10.1056/NEJMoa1211103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 202.Hamerman JA, Jarjoura JR, Humphrey MB, Nakamura MC, Seaman WE, Lanier LL. Cutting edge: inhibition of TLR and FcR responses in macrophages by triggering receptor expressed on myeloid cells (TREM)-2 and DAP12. J Immunol. 2006;177:2051–2055. doi: 10.4049/jimmunol.177.4.2051. [DOI] [PubMed] [Google Scholar]
  • 203.Turnbull IR, Gilfillan S, Cella M, Aoshi T, Miller M, Piccio L, Hernandez M, Colonna M. Cutting edge: TREM-2 attenuates macrophage activation. J Immunol. 2006;177:3520–3524. doi: 10.4049/jimmunol.177.6.3520. [DOI] [PubMed] [Google Scholar]
  • 204.Zhang B, Gaiteri C, Bodea LG, Wang Z, McElwee J, Podtelezhnikov AA, Zhang C, Xie T, Tran L, Dobrin R, Fluder E, Clurman B, Melquist S, Narayanan M, Suver C, Shah H, Mahajan M, Gillis T, Mysore J, MacDonald ME, Lamb JR, Bennett DA, Molony C, Stone DJ, Gudnason V, Myers AJ, Schadt EE, Neumann H, Zhu J, Emilsson V. Integrated systems approach identifies genetic nodes and networks in late-onset Alzheimer’s disease. Cell. 2013;153:707–720. doi: 10.1016/j.cell.2013.03.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 205.Moehle MS, Webber PJ, Tse T, Sukar N, Standaert DG, DeSilva TM, Cowell RM, West AB. LRRK2 inhibition attenuates microglial inflammatory responses. J Neurosci. 2012;32:1602–1611. doi: 10.1523/JNEUROSCI.5601-11.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 206.Kim B, Yang MS, Choi D, Kim JH, Kim HS, Seol W, Choi S, Jou I, Kim EY, Joe EH. Impaired inflammatory responses in murine Lrrk2-knockdown brain microglia. PLoS One. 2012;7:e34693. doi: 10.1371/journal.pone.0034693. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 207.Gardet A, Benita Y, Li C, Sands BE, Ballester I, Stevens C, Korzenik JR, Rioux JD, Daly MJ, Xavier RJ, Podolsky DK. LRRK2 Is Involved in the IFN-γ Response and Host Response to Pathogens. J Immunol. 2010;185:5577–5585. doi: 10.4049/jimmunol.1000548. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 208.Jun G, Naj AC, Beecham GW, Wang LS, Buros J, Gallins PJ, Buxbaum JD, Ertekin-Taner N, Fallin MD, Friedland R, Inzelberg R, Kramer P, Rogaeva E, St George-Hyslop P, Arnold SE, Baldwin CT, Barber R, Beach T, Bigio EH, Bird TD, Boxer A, Burke JR, Cairns N, Carroll SL, Chui HC, Clark DG, Cotman CW, Cummings JL, DeCarli C, Diaz-Arrastia R, Dick M, Dickson DW, Ellis WG, Fallon KB, Farlow MR, Ferris S, Frosch MP, Galasko DR, Gearing M, Geschwind DH, Ghetti B, Gilman S, Giordani B, Glass J, Graff-Radford NR, Green RC, Growdon JH, Hamilton RL, Harrell LE, Head E, Honig LS, Hulette CM, Hyman BT, Jicha GA, Jin LW, Johnson N, Karlawish J, Karydas A, Kaye JA, Kim R, Koo EH, Kowall NW, Lah JJ, Levey AI, Lieberman A, Lopez OL, Mack WJ, Markesbery W, Marson DC, Martiniuk F, Masliah E, McKee AC, Mesulam M, Miller JW, Miller BL, Miller CA, Parisi JE, Perl DP, Peskind E, Petersen RC, Poon W, Quinn JF, Raskind M, Reisberg B, Ringman JM, Roberson ED, Rosenberg RN, Sano M, Schneider JA, Schneider LS, Seeley W, Shelanski ML, Smith CD, Spina S, Stern RA, Tanzi RE, Trojanowski JQ, Troncoso JC, Van Deerlin VM, Vinters HV, Vonsattel JP, Weintraub S, Welsh-Bohmer KA, Woltjer RL, Younkin SG, Cantwell LB, Dombroski BA, Saykin AJ, Reiman EM, Bennett DA, Morris JC, Lunetta KL, Martin ER, Montine TJ, Goate AM, Blacker D, Tsuang DW, Beekly D, Cupples LA, Hakonarson H, Kukull W, Foroud TM, Haines J, Mayeux R, Farrer LA, Pericak-Vance MA, Schellenberg GD Alzheimer’s Disease Genetics Consortium. Meta-analysis Confirms CR1, CLU, and PICALM as Alzheimer Disease Risk Loci and Reveals Interactions With APOE Genotypes. Arch Neurol. 2010;67:1473–1484. doi: 10.1001/archneurol.2010.201. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 209.Chibnik LB, Shulman JM, Leurgans SE, Schneider JA, Wilson RS, Tran D, Aubin C, Buchman AS, Heward CB, Myers AJ, Hardy JA, Huentelman MJ, Corneveaux JJ, Reiman EM, Evans DA, Bennett DA, De Jager PL. CR1 is associated with amyloid plaque burden and age-related cognitive decline. Ann Neurol. 2011;69:560–569. doi: 10.1002/ana.22277. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 210.Wijsman EM, Pankratz ND, Choi Y, Rothstein JH, Faber KM, Cheng R, Lee JH, Bird TD, Bennett DA, Diaz-Arrastia R, Goate AM, Farlow M, Ghetti B, Sweet RA, Foroud TM, Mayeux R NIA-LOAD/NCRAD Family Study Group. Genome-Wide Association of Familial Late-Onset Alzheimer’s Disease Replicates BIN1 and CLU and Nominates CUGBP2 in Interaction with APOE. PLoS Genet. 2011;7:e1001308. doi: 10.1371/journal.pgen.1001308. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 211.Barral S, Bird T, Goate A, Farlow MR, Diaz-Arrastia R, Bennett DA, Graff-Radford N, Boeve BF, Sweet RA, Stern Y, Wilson RS, Foroud T, Ott J, Mayeux R National Institute on Aging Late-Onset Alzheimer’s Disease Genetics Study. Genotype patterns at PICALM, CR1, BIN1, CLU, and APOE genes are associated with episodic memory. Neurology. 2012;78:1464–1471. doi: 10.1212/WNL.0b013e3182553c48. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 212.Keenan BT, Shulman JM, Chibnik LB, Raj T, Tran D, Sabuncu MR Alzheimer’s Disease Neuroimaging Initiative. Allen AN, Corneveaux JJ, Hardy JA, Huentelman MJ, Lemere CA, Myers AJ, Nicholson-Weller A, Reiman EM, Evans DA, Bennett DA, De Jager PL. A coding variant in CR1 interacts with APOE-epsilon4 to influence cognitive decline. Hum Mol Genet. 2012;21:2377–2388. doi: 10.1093/hmg/dds054. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 213.Stephan AH, Barres BA, Stevens B. The complement system: an unexpected role in synaptic pruning during development and disease. Annu Rev Neurosci. 2012;35:369–389. doi: 10.1146/annurev-neuro-061010-113810. [DOI] [PubMed] [Google Scholar]
  • 214.Evans MC, Couch Y, Sibson N, Turner MR. Inflammation and neurovascular changes in amyotrophic lateral sclerosis. Mol Cell Neurosci. 2013;53:34–41. doi: 10.1016/j.mcn.2012.10.008. [DOI] [PubMed] [Google Scholar]
  • 215.Yin F, Banerjee R, Thomas B, Zhou P, Qian L, Jia T, Ma X, Ma Y, Iadecola C, Beal MF, Nathan C, Ding A. Exaggerated inflammation, impaired host defense, and neuropathology in progranulin-deficient mice. J Exp Med. 2010;207:117–128. doi: 10.1084/jem.20091568. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216.Jones KA, Thomsen C. The role of the innate immune system in psychiatric disorders. Mol Cell Neurosci. 2013;53:52–62. doi: 10.1016/j.mcn.2012.10.002. [DOI] [PubMed] [Google Scholar]
  • 217.Vom Berg J, Prokop S, Miller KR, Obst J, Kalin RE, Lopategui-Cabezas I, Wegner A, Mair F, Schipke CG, Peters O, Winter Y, Becher B, Heppner FL. Inhibition of IL-12/IL-23 signaling reduces Alzheimer’s disease-like pathology and cognitive decline. Nat Med. 2012;18:1812–1819. doi: 10.1038/nm.2965. [DOI] [PubMed] [Google Scholar]
  • 218.Gao HM, Liu B, Zhang W, Hong JS. Novel anti-inflammatory therapy for Parkinson’s disease. Trends Pharmacol Sci. 2003;24:395–401. doi: 10.1016/S0165-6147(03)00176-7. [DOI] [PubMed] [Google Scholar]
  • 219.Tan ZS, Seshadri S. Inflammation in the Alzheimer’s disease cascade: culprit or innocent bystander? Alzheimers Res Ther. 2010;2:6. doi: 10.1186/alzrt29. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 220.Martins IJ, Hone E, Foster JK, Sunram-Lea SI, Gnjec A, Fuller SJ, Nolan D, Gandy SE, Martins RN. Apolipoprotein E, cholesterol metabolism,diabetes, and the convergence of risk factors for Alzheimer’s disease and cardiovascular disease. Mol Psychiatry. 2006;11:721–736. doi: 10.1038/sj.mp.4001854. [DOI] [PubMed] [Google Scholar]
  • 221.Sparks DL. Alzheimer disease: statins in the treatment of Alzheimer disease. Nat Rev Neurol. 2011;7:662–663. doi: 10.1038/nrneurol.2011.165. [DOI] [PubMed] [Google Scholar]
  • 222.Gao X, Simon KC, Schwarzschild MA, Ascherio A. Prospective study of statin use and risk of Parkinson disease. Arch Neurol. 2012;69:380–384. doi: 10.1001/archneurol.2011.1060. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 223.Karasinska JM, Hayden MR. Cholesterol metabolism in Huntington disease. Nat Rev Neurol. 2011;7:561–572. doi: 10.1038/nrneurol.2011.132. [DOI] [PubMed] [Google Scholar]
  • 224.Anchisi L, Dessi S, Pani A, Mandas A. Cholesterol homeostasis: a key to prevent or slow down neurodegeneration. Front Physiol. 2012;3:486. doi: 10.3389/fphys.2012.00486. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 225.Han X, Rozen S, Boyle SH, Hellegers C, Cheng H, Burke JR, Welsh-Bohmer KA, Doraiswamy PM, Kaddurah-Daouk R. Metabolomics in early Alzheimer’s disease: identification of altered plasma sphingolipidome using shotgun lipidomics. PLoS One. 2011;6:e21643. doi: 10.1371/journal.pone.0021643. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 226.Wood PL. Lipidomics of Alzheimer’s disease: current status. Alzheimers Res Ther. 2012;4:5. doi: 10.1186/alzrt103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 227.Holtzman DM, Herz J, Bu G. Apolipoprotein e and apolipoprotein e receptors: normal biology and roles in Alzheimer disease. Cold Spring Harb Perspect Med. 2012;2:a006312. doi: 10.1101/cshperspect.a006312. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 228.Calero M, Tokuda T, Rostagno A, Kumar A, Zlokovic B, Frangione B, Ghiso J. Functional and structural properties of lipid-associated apolipoprotein J (clusterin) Biochem J. 1999;344:375–383. [PMC free article] [PubMed] [Google Scholar]
  • 229.Kim WS, Weickert CS, Garner B. Role of ATP-binding cassette transporters in brain lipid transport and neurological disease. J Neurochem. 2008;104:1145–1166. doi: 10.1111/j.1471-4159.2007.05099.x. [DOI] [PubMed] [Google Scholar]
  • 230.Soscia SJ, Fitzgerald ML. The ABCA7 transporter, brain lipids and Alzheimer’s disease. Clinical Lipidology. 2013;8:97–108. [Google Scholar]
  • 231.Biskup S, Moore DJ, Celsi F, Higashi S, West AB, Andrabi SA, Kurkinen K, Yu SW, Savitt JM, Waldvogel HJ, Faull RL, Emson PC, Torp R, Ottersen OP, Dawson TM, Dawson VL. Localization of LRRK2 to membranous and vesicular structures in mammalian brain. Ann Neurol. 2006;60:557–569. doi: 10.1002/ana.21019. [DOI] [PubMed] [Google Scholar]
  • 232.Kim KY, Stevens MV, Akter MH, Rusk SE, Huang RJ, Cohen A, Noguchi A, Springer D, Bocharov AV, Eggerman TL, Suen DF, Youle RJ, Amar M, Remaley AT, Sack MN. Parkin is a lipid-responsive regulator of fat uptake in mice and mutant human cells. J Clin Invest. 2011;121:3701–3712. doi: 10.1172/JCI44736. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 233.Hatano T, Kubo S, Imai S, Maeda M, Ishikawa K, Mizuno Y, Hattori N. Leucine-rich repeat kinase 2 associates with lipid rafts. Hum Mol Genet. 2007;16:678–690. doi: 10.1093/hmg/ddm013. [DOI] [PubMed] [Google Scholar]
  • 234.Chan RB, Oliveira TG, Cortes EP, Honig LS, Duff KE, Small SA, Wenk MR, Shui G, Di Paolo G. Comparative lipidomic analysis of mouse and human brain with Alzheimer disease. J Biol Chem. 2012;287:2678–2688. doi: 10.1074/jbc.M111.274142. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 235.Cheng D, Jenner AM, Shui G, Cheong WF, Mitchell TW, Nealon JR, Kim WS, McCann H, Wenk MR, Halliday GM, Garner B. Lipid pathway alterations in Parkinson’s disease primary visual cortex. PLoS One. 2011;6:e17299. doi: 10.1371/journal.pone.0017299. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 236.Ho PP, Kanter JL, Johnson AM, Srinagesh HK, Chang EJ, Purdy TM, van Haren K, Wikoff WR, Kind T, Khademi M, Matloff LY, Narayana S, Hur EM, Lindstrom TM, He Z, Fiehn O, Olsson T, Han X, Han MH, Steinman L, Robinson WH. Identification of Naturally Occurring Fatty Acids of the Myelin Sheath That Resolve Neuroinflammation. Sci Transl Med. 2012;4:137ra173. doi: 10.1126/scitranslmed.3003831. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 237.Willey JZ, Elkind MS. 3-Hydroxy-3-methylglutaryl-coenzyme A reductase inhibitors in the treatment of central nervous system diseases. Arch Neurol. 2010;67:1062–1067. doi: 10.1001/archneurol.2010.199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 238.Kroner Z. The relationship between Alzheimer’s disease and diabetes: type 3 diabetes? Altern Med Rev. 2009;14:373–379. [PubMed] [Google Scholar]
  • 239.Liu Y, Liu F, Grundke-Iqbal I, Iqbal K, Gong CX. Deficient brain insulin signalling pathway in Alzheimer’s disease and diabetes. J Pathol. 2011;225:54–62. doi: 10.1002/path.2912. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 240.Vignini A, Giulietti A, Nanetti L, Raffaelli F, Giusti L, Mazzanti L, Provinciali L. Alzheimer’s disease And Diabetes: New Insights and Unifying Therapies. Curr Diabetes Rev. 2013 May;9:218–27. doi: 10.2174/1573399811309030003. [DOI] [PubMed] [Google Scholar]
  • 241.de la Monte SM. Contributions of brain insulin resistance and deficiency in amyloid-related neurodegeneration in Alzheimer’s disease. Drugs. 2012;72:49–66. doi: 10.2165/11597760-000000000-00000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 242.Chen DY, Stern SA, Garcia-Osta A, Saunier-Rebori B, Pollonini G, Bambah-Mukku D, Blitzer RD, Alberini CM. A critical role for IGF-II in memory consolidation and enhancement. Nature. 2011;469:491–497. doi: 10.1038/nature09667. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 243.Lue LF, Andrade C, Sabbagh M, Walker D. Is There Inflammatory Synergy in Type II Diabetes Mellitus and Alzheimer’s Disease? Int J Alzheimers Dis. 2012;2012:918680. doi: 10.1155/2012/918680. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 244.Iwasaki T, Yoneda M, Nakajima A, Terauchi Y. Serum Butyrylcholinesterase is Strongly Associated with Adiposity, the Serum Lipid Profile and Insulin Resistance. Intern Med. 2007;46:1633–9. doi: 10.2169/internalmedicine.46.0049. [DOI] [PubMed] [Google Scholar]
  • 245.Meigs JB, Manning AK, Fox CS, Florez JC, Liu C, Cupples LA, Dupuis J. Genome-wide association with diabetes-related traits in the Framingham Heart Study. BMC Med Genet. 2007;8(Suppl 1):S16. doi: 10.1186/1471-2350-8-S1-S16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 246.Gao C, Holscher C, Liu Y, Li L. GSK3: a key target for the development of novel treatments for type 2 diabetes mellitus and Alzheimer disease. Rev Neurosci. 2012;23:1–11. doi: 10.1515/rns.2011.061. [DOI] [PubMed] [Google Scholar]
  • 247.Akinyemi R, Mukaetova-Ladinska E, Attems J, Attems J, Ihara M, Kalaria RN. Vascular Risk Factors and Neurodegeneration in Ageing related Dementias: Alzheimer’s Disease and Vascular Dementia. Curr Alzheimer Res. 2013 Jul;10:642–53. doi: 10.2174/15672050113109990037. [DOI] [PubMed] [Google Scholar]
  • 248.McCarron MO, Delong D, Alberts MJ. APOE genotype as a risk factor for ischemic cerebrovascular disease: a meta-analysis. Neurology. 1999;53:1308–1311. doi: 10.1212/wnl.53.6.1308. [DOI] [PubMed] [Google Scholar]
  • 249.Desai BS, Monahan AJ, Carvey PM, Hendey B. Blood-brain barrier pathology in Alzheimer’s and Parkinson’s disease: implications for drug therapy. Cell Transplant. 2007;16:285–299. doi: 10.3727/000000007783464731. [DOI] [PubMed] [Google Scholar]
  • 250.Chen X, Ghribi O, Geiger JD. Caffeine protects against disruptions of the blood-brain barrier in animal models of Alzheimer’s and Parkinson’s diseases. J Alzheimers Dis. 2010;20(Suppl 1):S127–141. doi: 10.3233/JAD-2010-1376. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 251.Bartels AL. Blood-brain barrier P-glycoprotein function in neurodegenerative disease. Curr Pharm Des. 2011;17:2771–2777. doi: 10.2174/138161211797440122. [DOI] [PubMed] [Google Scholar]
  • 252.Silver M, Janousova E, Hua X, Thompson PM, Montana G. Identification of gene pathways implicated in Alzheimer’s disease using longitudinal imaging phenotypes with sparse regression. Neuroimage. 2012;63:1681–1694. doi: 10.1016/j.neuroimage.2012.08.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 253.Biffi A, Shulman JM, Jagiella JM, Cortellini L, Ayres AM, Schwab K, Brown DL, Silliman SL, Selim M, Worrall BB, Meschia JF, Slowik A, De Jager PL, Greenberg SM, Schneider JA, Bennett DA, Rosand J. Genetic variation at CR1 increases risk of cerebral amyloid angiopathy. Neurology. 2012;78:334–341. doi: 10.1212/WNL.0b013e3182452b40. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 254.Mihci E, Ozkaynak SS, Sallakci N, Kizilay F, Yavuzer U. VEGF polymorphisms and serum VEGF levels in Parkinson’s disease. Neurosci Lett. 2011;494:1–5. doi: 10.1016/j.neulet.2011.02.027. [DOI] [PubMed] [Google Scholar]
  • 255.He D, Lu W, Chang K, Liu Y, Zhang J, Zeng Z. Vascular endothelial growth factor polymorphisms and risk of Alzheimer’s disease: a meta-analysis. Gene. 2013;518:296–302. doi: 10.1016/j.gene.2013.01.021. [DOI] [PubMed] [Google Scholar]
  • 256.Topiwala A, Ebmeier KP. Vascular changes and brain plasticity: a new approach to neurodegenerative diseases. Am J Neurodegener Dis. 2012;1:152–159. [PMC free article] [PubMed] [Google Scholar]
  • 257.Alvarez JI, Dodelet-Devillers A, Kebir H, Ifergan I, Fabre PJ, Terouz S, Sabbagh M, Wosik K, Bourbonniere L, Bernard M, van Horssen J, de Vries HE, Charron F, Prat A. The Hedgehog pathway promotes blood-brain barrier integrity and CNS immune quiescence. Science. 2011;334:1727–1731. doi: 10.1126/science.1206936. [DOI] [PubMed] [Google Scholar]
  • 258.Schrijvers EM, Schurmann B, Koudstaal PJ, van den Bussche H, Van Duijn CM, Hentschel F, Heun R, Hofman A, Jessen F, Kolsch H, Kornhuber J, Peters O, Rivadeneira F, Ruther E, Uitterlinden AG, Riedel-Heller S, Dichgans M, Wiltfang J, Maier W, Breteler MM, Ikram MA. Genome-wide association study of vascular dementia. Stroke. 2012;43:315–319. doi: 10.1161/STROKEAHA.111.628768. [DOI] [PubMed] [Google Scholar]
  • 259.Chuang DM, Leng Y, Marinova Z, Kim HJ, Chiu CT. Multiple roles of HDAC inhibition in neurodegenerative conditions. Trends Neurosci. 2009;32:591–601. doi: 10.1016/j.tins.2009.06.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 260.Egger G, Liang G, Aparicio A, Jones PA. Epigenetics in human disease and prospects for epigenetic therapy. Nature. 2004;429:457–463. doi: 10.1038/nature02625. [DOI] [PubMed] [Google Scholar]
  • 261.Fraga MF, Ballestar E, Paz MF, Ropero S, Setien F, Ballestar ML, Heine-Suner D, Cigudosa JC, Urioste M, Benitez J, Boix-Chornet M, Sanchez-Aguilera A, Ling C, Carlsson E, Poulsen P, Vaag A, Stephan Z, Spector TD, Wu YZ, Plass C, Esteller M. Epigenetic differences arise during the lifetime of monozygotic twins. Proc Natl Acad Sci U S A. 2005;102:10604–10609. doi: 10.1073/pnas.0500398102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 262.Kontopoulos E, Parvin JD, Feany MB. Alpha-synuclein acts in the nucleus to inhibit histone acetylation and promote neurotoxicity. Hum Mol Genet. 2006;15:3012–3023. doi: 10.1093/hmg/ddl243. [DOI] [PubMed] [Google Scholar]
  • 263.Outeiro TF, Kontopoulos E, Altmann SM, Kufareva I, Strathearn KE, Amore AM, Volk CB, Maxwell MM, Rochet JC, McLean PJ, Young AB, Abagyan R, Feany MB, Hyman BT, Kazantsev AG. Sirtuin 2 inhibitors rescue alpha-synuclein-mediated toxicity in models of Parkinson’s disease. Science. 2007;317:516–519. doi: 10.1126/science.1143780. [DOI] [PubMed] [Google Scholar]
  • 264.Graff J, Rei D, Guan JS, Wang WY, Seo J, Hennig KM, Nieland TJ, Fass DM, Kao PF, Kahn M, Su SC, Samiei A, Joseph N, Haggarty SJ, Delalle I, Tsai LH. An epigenetic blockade of cognitive functions in the neurodegenerating brain. Nature. 2012;483:222–226. doi: 10.1038/nature10849. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 265.Donmez G, Wang D, Cohen DE, Guarente L. SIRT1 suppresses beta-amyloid production by activating the alpha-secretase gene ADAM10. Cell. 2010;142:320–332. doi: 10.1016/j.cell.2010.06.020. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  • 266.Beck J, Poulter M, Hensman D, Rohrer JD, Mahoney CJ, Adamson G, Campbell T, Uphill J, Borg A, Fratta P, Orrell RW, Malaspina A, Rowe J, Brown J, Hodges J, Sidle K, Polke JM, Houlden H, Schott JM, Fox NC, Rossor MN, Tabrizi SJ, Isaacs AM, Hardy J, Warren JD, Collinge J, Mead S. Large C9orf72 hexanucleotide repeat expansions are seen in multiple neurodegenerative syndromes and are more frequent than expected in the UK population. Am J Hum Genet. 2013;92:345–353. doi: 10.1016/j.ajhg.2013.01.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 267.Mori K, Weng SM, Arzberger T, May S, Rentzsch K, Kremmer E, Schmid B, Kretzschmar HA, Cruts M, Van Broeckhoven C, Haass C, Edbauer D. The C9orf72 GGGGCC repeat is translated into aggregating dipeptide-repeat proteins in FTLD/ALS. Science. 2013;339:1335–1338. doi: 10.1126/science.1232927. [DOI] [PubMed] [Google Scholar]
  • 268.Taylor JP. Neuroscience. RNA that gets RAN in neurodegeneration. Science. 2013;339:1282–1283. doi: 10.1126/science.1236450. [DOI] [PubMed] [Google Scholar]
  • 269.Rao JS, Keleshian VL, Klein S, Rapoport SI. Epigenetic modifications in frontal cortex from Alzheimer’s disease and bipolar disorder patients. Transl Psychiatry. 2012;2:e132. doi: 10.1038/tp.2012.55. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 270.Chowdary PD, Che DL, Cui B. Neurotrophin signaling via long-distance axonal transport. Annu Rev Phys Chem. 2012;63:571–594. doi: 10.1146/annurev-physchem-032511-143704. [DOI] [PubMed] [Google Scholar]
  • 271.Schindowski K, Belarbi K, Buee L. Neurotrophic factors in Alzheimer’s disease: role of axonal transport. Genes Brain Behav. 2008;7(Suppl 1):43–56. doi: 10.1111/j.1601-183X.2007.00378.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 272.Kauwe JS, Wang J, Mayo K, Morris JC, Fagan AM, Holtzman DM, Goate AM. Alzheimer’s disease risk variants show association with cerebrospinal fluid amyloid beta. Neurogenetics. 2009;10:13–17. doi: 10.1007/s10048-008-0150-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 273.Karamohamed S, Latourelle JC, Racette BA, Perlmutter JS, Wooten GF, Lew M, Klein C, Shill H, Golbe LI, Mark MH, Guttman M, Nicholson G, Wilk JB, Saint-Hilaire M, DeStefano AL, Prakash R, Tobin S, Williamson J, Suchowersky O, Labell N, Growdon BN, Singer C, Watts R, Goldwurm S, Pezzoli G, Baker KB, Giroux ML, Pramstaller PP, Burn DJ, Chinnery P, Sherman S, Vieregge P, Litvan I, Gusella JF, Myers RH, Parsian A. BDNF genetic variants are associated with onset age of familial Parkinson disease: GenePD Study. Neurology. 2005;65:1823–1825. doi: 10.1212/01.wnl.0000187075.81589.fd. [DOI] [PubMed] [Google Scholar]
  • 274.Voineskos AN, Lerch JP, Felsky D, Shaikh S, Rajji TK, Miranda D, Lobaugh NJ, Mulsant BH, Pollock BG, Kennedy JL. The brain-derived neurotrophic factor Val66Met polymorphism and prediction of neural risk for Alzheimer disease. Arch Gen Psychiatry. 2011;68:198–206. doi: 10.1001/archgenpsychiatry.2010.194. [DOI] [PubMed] [Google Scholar]
  • 275.Chen DY, Bambah-Mukku D, Pollonini G, Alberini CM. Glucocorticoid receptors recruit the CaMKII alpha-BDNF-CREB pathways to mediate memory consolidation. Nat Neuroscience. 2012;15:1707–14. doi: 10.1038/nn.3266. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 276.Choi JM, Hong JH, Chae MJ, Ngyuen PH, Kang HS, Ma HI, Kim YJ. Analysis of mutations and the association between polymorphisms in the cerebral dopamine neurotrophic factor (CDNF) gene and Parkinson disease. Neurosci Lett. 2011;493:97–101. doi: 10.1016/j.neulet.2011.02.013. [DOI] [PubMed] [Google Scholar]
  • 277.Lin LF, Doherty DH, Lile JD, Bektesh S, Collins F. GDNF: a glial cell line-derived neurotrophic factor for midbrain dopaminergic neurons. Science. 1993;260:1130–1132. doi: 10.1126/science.8493557. [DOI] [PubMed] [Google Scholar]
  • 278.Gash DM, Zhang Z, Ovadia A, Cass WA, Yi A, Simmerman L, Russell D, Martin D, Lapchak PA, Collins F, Hoffer BJ, Gerhardt GA. Functional recovery in parkinsonian monkeys treated with GDNF. Nature. 1996;380:252–255. doi: 10.1038/380252a0. [DOI] [PubMed] [Google Scholar]
  • 279.Patani N, Jiang WG, Mokbel K. Brain-derived neurotrophic factor expression predicts adverse pathological & clinical outcomes in human breast cancer. Cancer Cell Int. 2011;11:23. doi: 10.1186/1475-2867-11-23. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 280.Cao L, During MJ. What is the brain-cancer connection? Annu Rev Neurosci. 2012;35:331–345. doi: 10.1146/annurev-neuro-062111-150546. [DOI] [PubMed] [Google Scholar]
  • 281.Calado RT, Young NS. Telomere diseases. N Engl J Med. 2009;361:2353–2365. doi: 10.1056/NEJMra0903373. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 282.Hadley EC, Lakatta EG, Morrison-Bogorad M, Warner HR, Hodes RJ. The future of aging therapies. Cell. 2005;120:557–567. doi: 10.1016/j.cell.2005.01.030. [DOI] [PubMed] [Google Scholar]
  • 283.Honig LS, Kang MS, Schupf N, Lee JH, Mayeux R. Association of shorter leukocyte telomere repeat length with dementia and mortality. Arch Neurol. 2012;69:1332–1339. doi: 10.1001/archneurol.2012.1541. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 284.Rolyan H, Scheffold A, Heinrich A, Begus-Nahrmann Y, Langkopf BH, Holter SM, Vogt-Weisenhorn DM, Liss B, Wurst W, Lie DC, Thal DR, Biber K, Rudolph KL. Telomere shortening reduces Alzheimer’s disease amyloid pathology in mice. Brain. 2011;134:2044–2056. doi: 10.1093/brain/awr133. [DOI] [PubMed] [Google Scholar]
  • 285.Wang H, Chen H, Gao X, McGrath M, Deer D, De Vivo I, Schwarzschild MA, Ascherio A. Telomere length and risk of Parkinson’s disease. Mov Disord. 2008;23:302–305. doi: 10.1002/mds.21867. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 286.Eerola J, Kananen L, Manninen K, Hellstrom O, Tienari PJ, Hovatta I. No evidence for shorter leukocyte telomere length in Parkinson’s disease patients. J Gerontol A Biol Sci Med Sci. 2010;65:1181–1184. doi: 10.1093/gerona/glq125. [DOI] [PubMed] [Google Scholar]
  • 287.Hudson G, Faini D, Stutt A, Eccles M, Robinson L, Burn DJ, Chinnery PF. No evidence of substantia nigra telomere shortening in Parkinson’s disease. Neurobiol Aging. 2011;32:2107, e2103–2105. doi: 10.1016/j.neurobiolaging.2011.05.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 288.Njajou OT, Blackburn EH, Pawlikowska L, Mangino M, Damcott CM, Kwok PY, Spector TD, Newman AB, Harris TB, Cummings SR, Cawthon RM, Shuldiner AR, Valdes AM, Hsueh WC. A common variant in the telomerase RNA component is associated with short telomere length. PLoS One. 2010;5:e13048. doi: 10.1371/journal.pone.0013048. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 289.Shen Q, Zhang Z, Yu L, Cao L, Zhou D, Kan M, Li B, Zhang D, He L, Liu Y. Common variants near TERC are associated with leukocyte telomere length in the Chinese Han population. Eur J Hum Genet. 2011;19:721–723. doi: 10.1038/ejhg.2011.4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 290.Kim S, Parks CG, Xu Z, Carswell G, DeRoo LA, Sandler DP, Taylor JA. Association between genetic variants in DNA and histone methylation and telomere length. PLoS One. 2012;7:e40504. doi: 10.1371/journal.pone.0040504. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 291.Deelen J, Uh HW, Monajemi R, van Heemst D, Thijssen PE, Bohringer S, van den Akker EB, de Craen AJ, Rivadeneira F, Uitterlinden AG, Westendorp RG, Goeman JJ, Slagboom PE, Houwing-Duistermaat JJ, Beekman M. Gene set analysis of GWAS data for human longevity highlights the relevance of the insulin/IGF-1 signaling and telomere maintenance pathways. Age (Dordr) 2013;35:235–249. doi: 10.1007/s11357-011-9340-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 292.Fossel M. Telomerase and the aging cell: implications for human health. JAMA. 1998;279:1732–1735. doi: 10.1001/jama.279.21.1732. [DOI] [PubMed] [Google Scholar]
  • 293.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;39:17–23. doi: 10.1038/ng1934. [DOI] [PubMed] [Google Scholar]
  • 294.Lill CM, Roehr JT, McQueen MB, Kavvoura FK, Bagade S, Schjeide BM, Schjeide LM, Meissner E, Zauft U, Allen NC, Liu T, Schilling M, Anderson KJ, Beecham G, Berg D, Biernacka JM, Brice A, DeStefano AL, Do CB, Eriksson N, Factor SA, Farrer MJ, Foroud T, Gasser T, Hamza T, Hardy JA, Heutink P, Hill-Burns EM, Klein C, Latourelle JC, Maraganore DM, Martin ER, Martinez M, Myers RH, Nalls MA, Pankratz N, Payami H, Satake W, Scott WK, Sharma M, Singleton AB, Stefansson K, Toda T, Tung JY, Vance J, Wood NW, Zabetian CP 23andMe Genetic Epidemiology of Parkinson’s Disease Consortium; International Parkinson’s Disease Genomics Consortium; Parkinson’s Disease GWAS Consortium; Wellcome Trust Case Control Consortium 2) Young P, Tanzi RE, Khoury MJ, Zipp F, Lehrach H, Ioannidis JP, Bertram L. Comprehensive research synopsis and systematic meta-analyses in Parkinson’s disease genetics: The PDGene database. PLoS Genet. 2012;8:e1002548. doi: 10.1371/journal.pgen.1002548. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 295.Maloney B, Ge YW, Petersen RC, Hardy J, Rogers JT, Perez-Tur J, Lahiri DK. Functional characterization of three single-nucleotide polymorphisms present in the human APOE promoter sequence: Differential effects in neuronal cells and on DNA-protein interactions. Am J Med Genet B Neuropsychiatr Genet. 2010;153B:185–201. doi: 10.1002/ajmg.b.30973. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 296.Citron BA, Dennis JS, Zeitlin RS, Echeverria V. Transcription factor Sp1 dysregulation in Alzheimer’s disease. J Neurosci Res. 2008;86:2499–2504. doi: 10.1002/jnr.21695. [DOI] [PubMed] [Google Scholar]
  • 297.Chu J, Zhuo JM, Pratico D. Transcriptional regulation of beta-secretase-1 by 12/15-lipoxygenase results in enhanced amyloidogenesis and cognitive impairments. Ann Neurol. 2012;71:57–67. doi: 10.1002/ana.22625. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 298.Santpere G, Nieto M, Puig B, Ferrer I. Abnormal Sp1 transcription factor expression in Alzheimer disease and tauopathies. Neurosci Lett. 2006;397:30–34. doi: 10.1016/j.neulet.2005.11.062. [DOI] [PubMed] [Google Scholar]
  • 299.Karin M, Liu Z, Zandi E. AP-1 function and regulation. Curr Opin Cell Biol. 1997;9:240–246. doi: 10.1016/s0955-0674(97)80068-3. [DOI] [PubMed] [Google Scholar]
  • 300.Calon F, Grondin R, Morissette M, Goulet M, Blanchet PJ, Di Paolo T, Bedard PJ. Molecular basis of levodopa-induced dyskinesias. Ann Neurol. 2000;47:S70–78. [PubMed] [Google Scholar]
  • 301.Andersson M, Konradi C, Cenci MA. cAMP response element-binding protein is required for dopamine-dependent gene expression in the intact but not the dopamine-denervated striatum. J Neurosci. 2001;21:9930–9943. doi: 10.1523/JNEUROSCI.21-24-09930.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 302.Jegga AG, Schneider L, Ouyang X, Zhang J. Systems biology of the autophagy-lysosomal pathway. Autophagy. 2011;7:477–489. doi: 10.4161/auto.7.5.14811. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 303.Fahrig T, Gerlach I, Horvath E. A synthetic derivative of the natural product rocaglaol is a potent inhibitor of cytokine-mediated signaling and shows neuroprotective activity in vitro and in animal models of Parkinson’s disease and traumatic brain injury. Mol Pharmacol. 2005;67:1544–1555. doi: 10.1124/mol.104.008177. [DOI] [PubMed] [Google Scholar]
  • 304.Perez-Cadahia B, Drobic B, Davie JR. Activation and function of immediate-early genes in the nervous system. Biochem Cell Biol. 2011;89:61–73. doi: 10.1139/O10-138. [DOI] [PubMed] [Google Scholar]
  • 305.Gomez Ravetti M, Rosso OA, Berretta R, Moscato P. Uncovering molecular biomarkers that correlate cognitive decline with the changes of hippocampus’ gene expression profiles in Alzheimer’s disease. PLoS One. 2010;5:e10153. doi: 10.1371/journal.pone.0010153. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 306.Lu Y, Li T, Qureshi HY, Han D, Paudel HK. Early growth response 1 (Egr-1) regulates phosphorylation of microtubule-associated protein tau in mammalian brain. J Biol Chem. 2011;286:20569–20581. doi: 10.1074/jbc.M111.220962. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 307.Lindersson EK, Hojrup P, Gai WP, Locker D, Martin D, Jensen PH. alpha-Synuclein filaments bind the transcriptional regulator HMGB-1. Neuroreport. 2004;15:2735–2739. [PubMed] [Google Scholar]
  • 308.Takata K, Kitamura Y, Kakimura J, Shibagaki K, Tsuchiya D, Taniguchi T, Smith MA, Perry G, Shimohama S. Role of high mobility group protein-1 (HMG1) in amyloid-beta homeostasis. Biochem Biophys Res Commun. 2003;301:699–703. doi: 10.1016/s0006-291x(03)00024-x. [DOI] [PubMed] [Google Scholar]
  • 309.Takata K, Takada T, Ito A, Asai M, Tawa M, Saito Y, Ashihara E, Tomimoto H, Kitamura Y, Shimohama S. Microglial Amyloid-beta1-40 Phagocytosis Dysfunction Is Caused by High-Mobility Group Box Protein-1: Implications for the Pathological Progression of Alzheimer’s Disease. Int J Alzheimers Dis. 2012;2012:685739. doi: 10.1155/2012/685739. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 310.Scaffidi P, Misteli T, Bianchi ME. Release of chromatin protein HMGB1 by necrotic cells triggers inflammation. Nature. 2002;418:191–195. doi: 10.1038/nature00858. [DOI] [PubMed] [Google Scholar]
  • 311.Gao HM, Zhou H, Zhang F, Wilson BC, Kam W, Hong JS. HMGB1 acts on microglia Mac1 to mediate chronic neuroinflammation that drives progressive neurodegeneration. J Neurosci. 2011;31:1081–1092. doi: 10.1523/JNEUROSCI.3732-10.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 312.Clifford DB, Fagan AM, Holtzman DM, Morris JC, Teshome M, Shah AR, Kauwe JS. CSF biomarkers of Alzheimer disease in HIV-associated neurologic disease. Neurology. 2009;73:1982–1987. doi: 10.1212/WNL.0b013e3181c5b445. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 313.Giunta B, Hou H, Zhu Y, Rrapo E, Tian J, Takashi M, Commins D, Singer E, He J, Fernandez F, Tan J. HIV-1 Tat contributes to Alzheimer’s disease-like pathology in PSAPP mice. Int J Clin Exp Pathol. 2009;2:433–443. [PMC free article] [PubMed] [Google Scholar]
  • 314.Hickman SE, Allison EK, El Khoury J. Microglial dysfunction and defective beta-amyloid clearance pathways in aging Alzheimer’s disease mice. J Neurosci. 2008;28:8354–8360. doi: 10.1523/JNEUROSCI.0616-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 315.Tarassishin L, Loudig O, Bauman A, Shafit-Zagardo B, Suh HS, Lee SC. Interferon regulatory factor 3 inhibits astrocyte inflammatory gene expression through suppression of the proinflammatory miR-155 and miR-155*. Glia. 2011;59:1911–1922. doi: 10.1002/glia.21233. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 316.Chakrabarty P, Ceballos-Diaz C, Lin WL, Beccard A, Jansen-West K, McFarland NR, Janus C, Dickson D, Das P, Golde TE. Interferon-gamma induces progressive nigrostriatal degeneration and basal ganglia calcification. Nat Neurosci. 2011;14:694–696. doi: 10.1038/nn.2829. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 317.Soares HD, Potter WZ, Pickering E, Kuhn M, Immermann FW, Shera DM, Ferm M, Dean RA, Simon AJ, Swenson F, Siuciak JA, Kaplow J, Thambisetty M, Zagouras P, Koroshetz WJ, Wan HI, Trojanowski JQ, Shaw LM Biomarkers Consortium Alzheimer’s Disease Plasma Proteomics Project. Plasma biomarkers associated with the apolipoprotein E genotype and Alzheimer disease. Arch Neurol. 2012;69:1310–1317. doi: 10.1001/archneurol.2012.1070. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 318.Barcia C, Ros CM, Annese V, Gomez A, Ros-Bernal F, Aguado-Yera D, Martinez-Pagan ME, de Pablos V, Fernandez-Villalba E, Herrero MT. IFN-gamma signaling, with the synergistic contribution of TNF-alpha, mediates cell specific microglial and astroglial activation in experimental models of Parkinson’s disease. Cell Death Dis. 2011;2:e142. doi: 10.1038/cddis.2011.17. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 319.Jaeger S, Pietrzik CU. Functional role of lipoprotein receptors in Alzheimer’s disease. Curr Alzheimer Res. 2008;5:15–25. doi: 10.2174/156720508783884675. [DOI] [PubMed] [Google Scholar]
  • 320.Waldron E, Heilig C, Schweitzer A, Nadella N, Jaeger S, Martin AM, Weggen S, Brix K, Pietrzik CU. LRP1 modulates APP trafficking along early compartments of the secretory pathway. Neurobiol Dis. 2008;31:188–197. doi: 10.1016/j.nbd.2008.04.006. [DOI] [PubMed] [Google Scholar]
  • 321.Liu Q, Trotter J, Zhang J, Peters MM, Cheng H, Bao J, Han X, Weeber EJ, Bu G. Neuronal LRP1 knockout in adult mice leads to impaired brain lipid metabolism and progressive, age-dependent synapse loss and neurodegeneration. J Neurosci. 2010;30:17068–17078. doi: 10.1523/JNEUROSCI.4067-10.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 322.Sloan CD, Shen L, West JD, Wishart HA, Flashman LA, Rabin LA, Santulli RB, Guerin SJ, Rhodes CH, Tsongalis GJ, McAllister TW, Ahles TA, Lee SL, Moore JH, Saykin AJ. Genetic pathway-based hierarchical clustering analysis of older adults with cognitive complaints and amnestic mild cognitive impairment using clinical and neuroimaging phenotypes. Am J Med Genet B Neuropsychiatr Genet. 2010;153B:1060–1069. doi: 10.1002/ajmg.b.31078. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 323.Manolio TA, Collins FS, Cox NJ, Goldstein DB, Hindorff LA, Hunter DJ, McCarthy MI, Ramos EM, Cardon LR, Chakravarti A, Cho JH, Guttmacher AE, Kong A, Kruglyak L, Mardis E, Rotimi CN, Slatkin M, Valle D, Whittemore AS, Boehnke M, Clark AG, Eichler EE, Gibson G, Haines JL, Mackay TF, McCarroll SA, Visscher PM. Finding the missing heritability of complex diseases. Nature. 2009;461:747–753. doi: 10.1038/nature08494. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 324.Zuk O, Hechter E, Sunyaev SR, Lander ES. The mystery of missing heritability: Genetic interactions create phantom heritability. Proc Natl Acad Sci U S A. 2012;109:1193–1198. doi: 10.1073/pnas.1119675109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 325.Peregrin-Alvarez JM, Sanford C, Parkinson J. The conservation and evolutionary modularity of metabolism. Genome Biol. 2009;10:R63. doi: 10.1186/gb-2009-10-6-r63. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 326.Ayalew M, Le-Niculescu H, Levey DF, Jain N, Changala B, Patel SD, Winiger E, Breier A, Shekhar A, Amdur R, Koller D, Nurnberger JI, Corvin A, Geyer M, Tsuang MT, Salomon D, Schork NJ, Fanous AH, O’Donovan MC, Niculescu AB. Convergent functional genomics of schizophrenia: from comprehensive understanding to genetic risk prediction. Mol Psychiatry. 2012;17:887–905. doi: 10.1038/mp.2012.37. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 327.Zhong H, Yang X, Kaplan LM, Molony C, Schadt EE. Integrating pathway analysis and genetics of gene expression for genome-wide association studies. Am J Hum Genet. 2010;86:581–591. doi: 10.1016/j.ajhg.2010.02.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 328.Ala-Korpela M, Kangas AJ, Inouye M. Genome-wide association studies and systems biology: together at last. Trends Genet. 2011;27:493–498. doi: 10.1016/j.tig.2011.09.002. [DOI] [PubMed] [Google Scholar]

Articles from American Journal of Neurodegenerative Disease are provided here courtesy of e-Century Publishing Corporation

RESOURCES