26
Views
25
CrossRef citations to date
0
Altmetric
Review

An update on the toxicity of Aβ in Alzheimer’s disease

, , &
Pages 1033-1042 | Published online: 05 Dec 2008

Abstract

Alzheimer’s disease is characterized histopathologically by deposition of insoluble forms of the peptide Aβ and the protein tau in brain. Aβ is the principal component of amyloid plaques and tau of neurofibrillary tangles. Familial cases of AD are associated with causal mutations in the gene encoding the amyloid precursor protein, APP, from which the amyloidogenic Aβ peptide is derived, and this supports a role for Aβ in disease. Aβ can promote tau pathology and at the same time its toxicity is also tau-dependent. Aβ can adopt different conformations including soluble oligomers and insoluble fibrillar species present in plaques. We discuss which of these conformations exert toxicity, highlight molecular pathways involved and discuss what has been learned by applying functional genomics.

Incidence of dementia

Alzheimer’s disease (AD) is the most common cause of dementia, comprising 50%–70% of all cases and affecting more than 15 million people worldwide. Frontotemporal dementia (FTD), in comparison, is less common but may make up to 50% of dementia cases presenting before the age of 60 (CitationGraff-Radford and Woodruff 2007). Dementia is defined as the significant loss of intellectual abilities such as memory functions, severe enough to interfere with social or occupational functioning. At present, AD cannot be cured which is different from memory dysfunction caused by malnutrition, drug abuse or depression where some form of treatment is available (CitationPatel et al 2007).

Neuropathology of AD and FTD

The AD brain is characterized by massive neuronal cell and synapse loss at specific predilection sites (CitationSelkoe 2002). The extracellular plaques and the intracellular neurofibrillary tangles are the key histopathological hallmarks of AD. The major proteinaceous component of the amyloid plaques is a 40- to 42-amino acid polypeptide termed Aβ (Aβ40 and Aβ42), which is derived by proteolytic cleavage from the amyloid precursor protein, APP, as part of normal cellular metabolism (CitationGlenner and Wong 1984; CitationMasters et al 1985). β-Secretase is the protease that generates the amino terminus of Aβ and γ-secretase cleavage at the carboxy-terminus dictates its length. Aβ40 is the most common species and Aβ42 is the more fibrillogenic and neurotoxic species. Recent evidence suggests that Aβ40 may prevent Aβ42 from aggregating and forming plaques (CitationYan and Wang 2007). β-Secretase activity has been attributed to a single protein, BACE 1 (CitationVassar et al 1999), whereas γ-secretase activity depends on four components, presenilin, nicastrin, APH-1 and PEN-2 forming a proteolytic complex (CitationEdbauer et al 2003). α-Secretase is involved in the non-amyloidogenic pathway by cleaving APP within the Aβ domain, thus precluding Aβ formation (CitationGotz and Ittner 2008). Which higher order forms of Aβ exert toxicity is a matter of debate. The conflicting data as well as putative mechanisms of toxicity are discussed in detail below.

The second histopathological hallmark of AD are the neurofibrillary lesions that are found in cell bodies and apical dendrites as neurofibrillary tangles, in distal dendrites as neuropil threads, and in the abnormal neurites that are associated with some β-amyloid plaques (neuritic plaques). Neurofibrillary tangles are also abundant, in the absence of overt plaques, in FTD and other so-called tauopathies (CitationLee et al 2001). The neurofibrillary lesions contain aggregates of the microtubule-associated protein tau that under physiological conditions is mainly localized to the axonal compartment of neurons (CitationGoedert et al 1988). In tauopathies such as PSP (Progressive Supranuclear Palsy) or CBD (Corticobasal Degeneration), tau also forms aggregates in non-neuronal cells (CitationGotz 2001), emphasizing the important role of glia in disease (CitationKurosinski and Gotz 2002). Tau has an unusually high content of serine and threonine residues and many of these are phosphorylated under physiological conditions (CitationChen et al 2004a). Under pathological conditions, tau is hyperphosphorylated, which means that it is phosphorylated to a higher degree at physiological sites, and at additional “pathological” sites. Phosphorylation tends to dissociate tau from microtubules. Tau also undergoes a conformational change which likely assists in differential phosphorylation (CitationJicha et al 1997). Both tau and Aβ undergo nucleation-dependent fibril formation (CitationHarper and Lansbury 1997). In the course of this process, initially dispersed polypeptide chains slowly come together to form a diverse array of fibrillation nuclei that enable the rapid outgrowth into higher order assemblies including fibrils (CitationHortschansky et al 2005; CitationPellarin and Caflisch 2006; CitationGotz et al 2008).

Genetics of AD and FTD

In familial AD, autosomal dominant mutations have been identified in three genes: APP and the presenilin 1 (PS1) and presenilin 2 (PS2) genes. Together, they account for less than 1% of the total number of AD cases (CitationDelacourte et al 2002). In addition, several susceptibility genes have been identified but only the apolipoprotein E (APOE) gene has been unanimously confirmed and found to be associated with sporadic AD (CitationRocchi et al 2003). Clinically and histopathologically, early-onset familial AD cannot be discriminated from late-onset sporadic AD (CitationGotz 2001).

Whereas in AD no mutations were found in the MAPT gene encoding tau, they were identified in FTD with Parkinsonism linked to chromosome 17 (FTDP-17) (CitationHutton et al 1998; CitationPoorkaj et al 1998; CitationSpillantini et al 1998). This established that dysfunction of tau in itself can cause neurodegeneration and lead to dementia. The existence of a subgroup of FTD with no tau aggregation was enigmatic for some time leading to the coining of terms such as ‘dementia lacking distinctive histology’. This dementia with tau-negative and ubiquitin-positive lesions today is termed FTLD-U or FTDU-17 although this is misleading as it implies that the tau lesions in FTDP-17 are ubiquitin-negative which is not the case.

FTDU-17 is caused by loss-of-function mutations in the gene encoding progranulin (PGRN), a growth factor involved in multiple physiological and pathological processes including tumorigenesis (CitationBaker et al 2006; CitationCruts et al 2006). The TAR DNA-binding protein of 43 kDa (TDP-43) is a constituent of the ubiquitin-positive inclusions in both FTLD-U and sporadic amyotrophic lateral sclerosis (ALS) arguing for an overlap in the pathology between these two entities (CitationNeumann et al 2006). Similar to tau, in diseased brain, TDP-43 becomes hyperphosphorylated, ubiquitinated, and carboxy-terminally truncated. Mutations in the gene encoding valosin-containing protein cause frontotemporal dementia with inclusion body myopathy and Paget disease of bone (IBMPFD), a rare, autosomal-dominant disorder. As TDP-43, but not valosin-containing protein, is accumulating in the ubiquitin-positive inclusions in IBMPFD this would argue that valosin-containing protein gene mutations lead to a dominant negative loss or alteration of valosin-containing protein function culminating in impaired degradation of TDP-43. In other words, TDP-43 is a common pathologic substrate linking a variety of distinct patterns of FTLD-U pathology caused by different genetic alterations (CitationNeumann et al 2007).

Clinical features of AD and FTD

AD is characterized by early memory deficits, followed by a gradual erosion of other cognitive functions such as judgment, verbal fluency and orientation. The most severe neuropathological changes occur in the hippocampal formation, followed by the association cortices and subcortical structures, including the amygdala and the nucleus basalis of Meynert (CitationArnold et al 1991). Neurofibrillary tangles develop and spread in a predictable manner across the brain providing the basis for distinguishing six stages of disease progression: the transentorhinal Braak stages I–II represent clinically silent cases; the limbic stages III–IV incipient AD; and the neocortical stages V–VI fully developed AD. By using phosphorylation-dependent anti-tau antibodies such as AT8, neuronal changes can be visualized well before the actual formation of neurofibrillary tangles (CitationBraak and Braak 1991, Citation1995).

In contrast to AD, which is characterized predominantly by memory loss, FTD is mainly initiated with behavioral impairment. The average age of diagnosis is about 60, which is around 10 years before the average sporadic AD patient is diagnosed (CitationSnowden et al 2001; CitationWeder et al 2007). Patients may have an often asymmetrical atrophy of the frontal and temporal cortex. There is evidence that motor neuron disease and FTD coexist, and the motor symptoms may precede, coincide, or follow the development of cognitive and behavioral changes (CitationGraff-Radford and Woodruff 2007). In a significant subset of FTD, late parkinsonism is found (CitationLee et al 2001).

Animal models of AD and FTD

To better understand the role of Aβ and tau in AD and related disorders, experimental animal models have been developed which reproduce aspects of the neuropathological characteristics of these diseases (CitationGotz et al 2007; CitationGotz and Ittner 2008).

In 1995, Games and coworkers established the first Aβ plaque-forming mouse model by targeting high levels of the disease-linked V717F mutant form of APP in brain, using the platelet-derived growth factor mini-promoter for expression. These PDAPP mice showed many of the pathological features of AD, including extensive deposition of extracellular amyloid plaques, astrocytosis and neuritic dystrophy (CitationGames et al 1995). Similar features were observed in a second transgenic model established by Hsiao and coworkers by expressing the APPsw mutation inserted into a hamster prion protein cosmid vector (CitationHsiao et al 1996). The APP23 strain was established by expressing APPsw under the control of the neuronal mThy1.2 promoter, with a seven-fold overexpression of APP (CitationSturchler-Pierrat et al 1997; CitationStalder et al 1999). Subsequently, many more models have been developed such as the TgCRND8 or J20 mice (CitationJanus et al 2000; CitationMucke et al 2000). These mouse models have been instrumental in addressing aspects of Aβ toxicity and testing therapies such as vaccination trials (CitationGotz 2001; CitationGotz et al 2004b; CitationKulic et al 2006).

The first tau transgenic model was established by us in 1995, expressing the longest human brain tau isoform, without a pathogenic mutation, in mice using the hThy1 promoter for neuronal expression (CitationGotz et al 1995). Despite the lack of a neurofibrillary pathology, these mice modeled aspects of human AD, such as the somatodendritic localization of hyperphosphorylated tau and, therefore, represented an early ‘pre-tangle’ phenotype. The subsequent use of stronger promoters caused a more pronounced tau phenotype in transgenic mice (CitationIshihara et al 1999; CitationSpittaels et al 1999; CitationProbst et al 2000; CitationGotz and Nitsch 2001). Once the first pathogenic FTDP-17 mutations were identified in the MAPT gene in 1998, several groups expressed them in mice and achieved neurofibrillary tangle formation, both in neurons and in glial cells (CitationGotz and Ittner 2008). We, for example, expressed G272V and P301L mutant tau and obtained mice with aggregated tau and neurofibrillary tangles (CitationGotz et al 2001a; CitationGotz et al 2001b; CitationGotz et al 2001c; CitationDeters et al 2008). The P301L tau-expressing pR5 mice showed a behavioural impairment in amygdala- and hippocampus-dependent tasks; aspects of the behavioral impairment could be correlated with the aggregation pattern of the transgene (CitationPennanen et al 2004; CitationPennanen et al 2006). K369I transgenic mice, on the other hand, model Parkinsonism in FTD, in parts owing to expression of the transgene in the substantia nigra (CitationIttner et al 2008).

Cross-talk between Aβ and tau

The amyloid cascade hypothesis claims, in simplistic terms, that in the pathogenic cascade of AD, Aβ is upstream of tau (CitationHardy and Selkoe 2002). To address the interaction between Aβ and tau (CitationGotz et al 2004a), Aβ plaque-forming Tg2576 mice were crossed with tangle-forming P301L tau-transgenic JNPL3 mice; also, P301L tau transgenic pR5 mice were intracerebrally injected with fibrillar preparations of Aβ42 (CitationGotz et al 2001b; CitationLewis et al 2001). Both strategies caused an increased tau phosphorylation at pathological epitopes and neurofibrillary tangle formation, establishing a link between Aβ and tau in vivo (CitationGotz et al 2001b; CitationLewis et al 2001). Similarly, tangle formation was aggravated by infusing brain extracts of aged plaque-forming APP23 mice intracerebrally in P301L tau transgenic mice or by crossing APP23 and P301L tau transgenic mice (CitationBolmont et al 2007). Together, these studies established that Aβ exaggerates a pre-existing tau pathology supporting, at least in parts, the amyloid cascade hypothesis in mice.

Interestingly, recent evidence suggests that Aβ toxicity is also tau-dependent (CitationRoberson et al 2007). Reducing endogenous tau levels prevented behavioral deficits as assessed in the Morris water maze, without altering Aβ levels. This was achieved by crossing plaque-forming APP transgenic mice onto hetero- and homozygous tau knockout backgrounds (CitationRoberson et al 2007). Tau reduction also protected both transgenic and non-transgenic mice against pentylenetetrazole (PTZ)-mediated excitotoxicity as shown by dramatic changes in seizure severity and latency. We were able to reproduce these findings in a second plaque-forming APP transgenic mouse model (Ittner et al submitted). Earlier findings in cultured hippocampal neurons derived from tau knockout and transgenic mice supports the model that tau is required for Aβ toxicity (CitationRapoport et al 2002). Together, this suggests that a reduction in tau levels is a potentially powerful treatment strategy for AD and other neurological conditions that are associated with neurotoxicity. A mechanistic explanation is lacking and therefore the importance of this information should encourage more research groups to work on the interaction between Aβ and tau.

Functional genomics and Aβ toxicity

An unbiased approach to address the toxicity of tau and Aβ is by functional genomics, which encompasses transcriptomic and proteomic techniques (CitationChen et al 2004b; CitationHoerndli et al 2004; CitationDavid et al 2005a; CitationHoerndli et al 2005). These highlighted a role for oxidative stress (see below) and impaired axonal transport in disease (CitationGotz et al 2006). When we analyzed the proteome pattern of total brain from P301L tau transgenic pR5 and wild-type mice by proteomics, we discovered that it was mainly metabolic related proteins including mitochondrial respiratory chain complex components, antioxidant enzymes and synaptic proteins that were modified in P301L tau mice. Importantly, mitochondrial dysfunction could be functionally validated in the P301L tau mice. Furthermore, the reduction in mitochondrial complex V levels in the P301L tau mice was also decreased in human P301L FTDP-17 brains. Finally, P301L tau mitochondria displayed an increased vulnerability towards fibrillar preparations of Aβ42, suggesting a synergistic action of tau and Aβ pathology on the mitochondria (CitationDavid et al 2005b).

In a follow-up study we investigated the toxicity of oligomeric Aβ species that in recent years have been suggested to be the main culprit, rather than fibrillar Aβ (CitationHartley et al 1999; CitationNecula et al 2007). The identity of the Aβ toxic species in both AD brains and in experimentally generated systems correlates best with the soluble, rather than the insoluble fibrillar forms. However, at present, there is little chemical and structural detail at the molecular and atomic level about the various aggregates to properly define the toxic species (CitationCappai and Barnham 2008). Interestingly, in cortical brain cells obtained from P301L tau transgenic pR5 mice both oligomeric and fibrillar, but not monomeric Aβ42 caused a decreased mitochondrial membrane potential (CitationEckert et al 2008a). This was not observed with cerebellar preparations indicating selective vulnerability of cortical neurons. Furthermore, we measured reductions in state 3 respiration, the respiratory control ratio and uncoupled respiration when incubating P301L tau mitochondria either with oligomeric or fibrillar preparations of Aβ42. We also found that aging specifically increased the sensitivity of mitochondria to oligomeric Aβ42 damage indicating that while oligomeric and fibrillar Aβ42 are both toxic, they exert different degrees of toxicity in mitochondria from older animals (CitationEckert et al 2008a).

To understand which processes are disrupted by Aβ42 in the presence of tau aggregates a comparative proteomics study was performed in both a cellular and an in vivo system (CitationDavid et al 2006). P301L tau expressing neuroblastoma cells were treated with Aβ42 as prior studies had shown that this caused tau filament formation (CitationFerrari et al 2003; CitationPennanen and Gotz 2005). In parallel, the amygdala of P301L tau transgenic mice was stereotaxically injected with Aβ42 as this causes increased tangle formation (CitationGotz et al 2001b). When the deregulated proteins in the two experimental paradigms were classified, it was found that a significant fraction of the altered proteins belonged to the same functional categories, ie, stress response and metabolism. We also identified model-specific effects of Aβ42 treatment such as differences in cell signaling proteins in the cellular model and changes in cytoskeletal and synapse proteins in the amygdala. By Western blotting and immunohistochemistry, we were able to show that 72% of the tested candidates were altered in human AD brain with a major emphasis on stress-related unfolded protein responsive candidates. This highlights these processes as important initiators in the Aβ42-mediated pathogenic cascade in AD and further supports the role of unfolded proteins in the course of AD (CitationDavid et al 2006).

Human SH-SY5Y neuroblastoma cells were also investigated by transcriptomics to assess the role of P301L mutant tau expression and treatment with or without Aβ on gene regulation. We found that Aβ and P301L tau expression independently affected the regulation of genes controlling cell proliferation and synaptic elements. Moreover, Aβ and P301L tau acted synergistically on cell cycle and DNA damage genes, yet influenced specific genes within these categories. By using neuronally differentiated P301L tau cells, it was shown that Aβ treatment induced an early up-regulation of cell cycle control and synaptic genes. Together, the study showed that Aβ treatment and human tau over-expression in a cell culture model acted synergistically to promote aberrant cell cycle re-entry supporting the mitosis failure hypothesis in AD (CitationArendt 2003; CitationHoerndli et al 2007).

The Yin and Yang of Aβ

AD has been termed a synaptic failure (CitationSelkoe 2002). While Aβ can kill neurons, it can also act by causing synaptotoxicity which may be more relevant for the earlier stages of AD that are best characterized by synaptic loss rather than neuronal death. Loss of synaptic terminals or dendritic spines could cause the associated decline in cognitive functions that characterizes AD. Whether the neurotoxic and synaptotoxic actions of Aβ are separate activities or whether they share common mechanisms is not known (CitationCappai and Barnham 2008).

How cells respond to Aβ varies depending upon the concentration of Aβ used, which adds another level of complexity. While Aβ peptide added at micromolar concentrations to primary neuronal cultures induces cell death (CitationYankner et al 1990), low, subnanomolar concentrations are neurotrophic arguing in favor of a physiological function of Aβ (CitationYankner et al 1990). As discussed above, the neurotoxic activity of Aβ is dependent upon its aggregation state. When Aβ aggregation was induced this increased its neurotoxic activity suggesting that the toxic species was associated with the formation of fibrils (CitationPike et al 1991a; CitationPike et al 1991b; CitationPike et al 1993). At present, however, there is a major research focus on the role of non-fibrillar soluble Aβ as the toxic species in AD (CitationLambert et al 1998; CitationWalsh et al 2002; CitationSmith et al 2007). These species have been given different names, including Aβ-derived diffusible ligands (ADDLs) (CitationLambert et al 1998). globulomers (CitationBarghorn et al 2005) and the Aβ star species 56 (Aβ*56) (CitationLesne et al 2006). To assist in identifying these species, conformational antibodies have been developed that not only stabilize the Aβ protofibrils but also prevent mature amyloid fibril formation (CitationHabicht et al 2007).

Aβ can inhibit long-term potentiation (LTP), a model system for synaptic strengthening and memory (CitationLambert et al 1998; CitationWalsh et al 2002; CitationCleary et al 2005; CitationKlyubin et al 2005; CitationTrommer et al 2005). When cell medium containing abundant Aβ monomers and proposed oligomers, but not amyloid fibrils was microinjected into rat brain, this markedly inhibits hippocampal long-term potentiation (LTP) (CitationWalsh et al 2002). Immunodepletion from the medium of all Aβ species completely abrogated this effect. Pretreatment of the medium with insulin-degrading enzyme, which degrades Aβ monomers but not oligomers, did not prevent the inhibition of LTP, indicating a role for Aβ oligomers. These were shown to disrupt synaptic plasticity in vivo at concentrations found in human brain and cerebrospinal fluid, in the absence of monomeric or fibrillar amyloid. When cells were treated with γ-secretase inhibitors at doses which prevented oligomer formation but allowed appreciable monomer production, this no longer disrupted LTP, indicating that synaptotoxic Aβ oligomers can be targeted therapeutically (CitationWalsh et al 2002; CitationWalsh et al 2005).

In Neuro-2A cells, oligomers were shown to induce a tenfold greater increase in neurotoxicity as compared to fibrils (CitationStine et al 2003). However, whereas LTP seems to be inhibited by oligomeric Aβ only and not fibrillar Aβ, in a different experimental paradigm, the two species seem to have both toxic, yet diverse effects (CitationWhite et al 2005). Using rat astrocyte cultures, oligomeric Aβ42 was shown to induce initial high levels of the pro-inflammatory molecule IL-1β that decreased over time, whereas fibrillar Aβ caused increased levels over time (CitationWhite et al 2005). It has been suggested that their neurotoxic activity is associated with dimeric and trimeric species however the exact composition of these higher molecular weight Aβ species has not been determined (CitationWalsh et al 2002; CitationCleary et al 2005) and remains a crucial point to definitively establish their molecular identity. Together, this shows that the relative role of the toxicity of monomeric compared to oligomeric compared to fibrillar species is far from being resolved.

Aβ and downstream signaling

What are the down-stream effectors of Aβ toxicity? Aβ may act via a plethora of pathways to induce synaptic and neuronal degeneration (CitationSmall et al 2001). Aβ’s anti-LTP activity can be modulated with antagonists to the p38 MAP kinase (CitationWang et al 2004) and the Jun NH2-terminal kinase (JNK) pathways (CitationMinogue et al 2003), both of which have also been implicated in tau phosphorylation (CitationKins et al 2001; CitationKins et al 2003). While inhibitors of p38, JNK, GSK-3β and phosphatidylinositol 3-kinase showed either no or only minor inhibition of Aβ oligomer-mediated cell death in mouse hippocampal slices, inhibitors of MAPK kinase kinase, which is upstream of the extracellular signal-regulated kinases, significantly inhibited Aβ-mediated neuronal death (CitationChong et al 2006).

Another interesting kinase is the non-receptor tyrosine kinase Fyn, as it links Aβ and tau. Fyn is a known interaction partner of tau (CitationLee et al 1998). Furthermore, Fyn is necessary for the toxicity of ADDLs (an oligomeric form of Aβ) as Fyn knockout neurons are resistant to ADDL-mediated neuronal cell death (CitationLambert et al 1998). Moreover, Fyn knockout mice display reduced synaptotoxicity without affecting aberrant sprouting, when crossed with APP transgenic mice (CitationChin et al 2004). Fyn has a role in modulating synaptic activity and plasticity, by phosphorylating the NMDA receptor (CitationBraithwaite et al 2006). This finding is consistent with the fact that Aβ oligomers alter the transport of the NMDA receptor by promoting its endocytosis and resulting in decreased NMDA receptor activity both in vitro and in APP transgenic mice (CitationSnyder et al 2005). Work in neuronal and astrocyte cultures further suggests that Aβ causes Ca2+-dependent oxidative stress by activating an astrocytic NADPH oxidase, with neuronal death following through a failure of antioxidant support (CitationAbramov et al 2004). Together, this suggests a fine-balanced network of molecular interactions (CitationCappai and Barnham 2008).

At present it is not understood whether Aβ acts via a receptor or whether membrane binding alone is sufficient (CitationCappai and Barnham 2008). If Aβ acts via a receptor, this receptor may have specificity for Aβ or it may bind proteins or peptides with shared amyloid properties. Work in primary cortical and hippocampal cultures treated with Aβ and human amylin, respectively, indicates that the latter may be the case, as rat amylin, which is not amyloidogenic, turns out not to be toxic (CitationLim et al 2008). Membrane interaction of Aβ can occur via its hydrophobic carboxy-terminal domain (CitationBhatia et al 2000; CitationAmbroggio et al 2005) or by electrostatic interactions mediated by the charged amino acids in the amino-terminal domain (CitationLau et al 2006). Aβ may bind to the cell membrane forming channels or pores that disrupt ion homeostasis, hence leading to neuronal dysfunction (CitationArispe et al 1993; CitationPollard et al 1995; CitationHolscher 1998; CitationBhatia et al 2000; CitationLin et al 2001). As several molecules associated with disease (such as the Prion protein, the British peptide, or human amylin) can form soluble oligomers and bind to membranes and disrupt ion homeostasis, this may be an inherent property of amyloidogenic proteins or peptides (CitationDemuro et al 2005).

Aβ binding proteins

A number of Aβ-binding proteins have been identified on the plasma membrane of neurons and glial cells. These include the alpha7 nicotinic acetylcholine receptor, the receptor for advanced glycosylation end-products (RAGE), APP itself, the NMDA receptor, the P75 neurotrophin receptor (P75NTR), the scavenger receptors, CD36 and low-density lipoprotein receptor-related protein (LRP) members (CitationVerdier et al 2004). RAGE can bind both non-fibrillar and fibrillar forms of Aβ (CitationYan et al 1996). Transgenic mice co-expressing mutant APP and RAGE revealed an earlier onset of memory defects and synaptic dysfunction than single APP transgenic mice (CitationArancio et al 2004). LRP, apoE and the serum protein α2-macroglobulin (α2M) probably modulate Aβ toxicity via clearance of apoE:Aβ and α2M:Aβ complexes or Aβ alone from the brain and hence reduce Aβ levels (CitationShibata et al 2000; CitationDeane et al 2004). The addition of an anti-NMDA receptor antibody can block Aβ oligomer binding to neurons and reduce ROS stimulation in hippocampal cultures (CitationDe Felice et al 2007) suggesting a direct interaction between these two proteins. Alternatively, Aβ may be interacting with NMDA receptor via an integrin-mediated effect (CitationBi et al 2002). P75NTR can bind a variety of Aβ oligomeric species and modulate Aβ toxicity in a cell type- and P75 isoform-dependent manner (CitationCoulson 2006; CitationSotthibundhu et al 2008). Full-length P75NTR blocks toxicity of fibrillar and non-fibrillar Aβ in primary neurons (CitationZhang et al 2003), but promotes toxicity of fibrillar Aβ in neuroblastoma cells (CitationPerini et al 2002).

Aβ may not only bind to the cell surface but also act on intracellular organelles such as mitochondria (CitationLustbader et al 2004; CitationCaspersen et al 2005; CitationCrouch et al 2005; CitationDevi et al 2006; CitationManczak et al 2006) whose function it impairs (CitationKeil et al 2004; CitationDavid et al 2005b; CitationEckert et al 2008a; CitationEckert et al 2008b). Mitochondrial dysfunction was also linked to full-length and carboxy-terminally truncated APP, that was shown to accumulate exclusively in the protein import channels of mitochondria of human AD, but not age-matched control brains (CitationDevi et al 2006). Similarly, accumulation of full-length APP in the mitochondrial compartment in a trans-membrane-arrested form, but not lacking the acidic domain, was shown to cause mitochondrial dysfunction and impair energy metabolism (CitationAnandatheerthavarada et al 2003). Aβ can disrupt mitochondrial cytochrome c oxidase activity (CitationCrouch et al 2005; CitationTakuma et al 2005) in a sequence- and conformer-dependent manner (CitationCrouch et al 2005). The Aβ binding protein alcohol dehydrogenase (ABAD) is a short-chain alcohol dehydrogenase that binds to Aβ in the mitochondrial matrix. This lead to mitochondrial failure via changes in mitochondrial membrane permeability and a reduction in the activities of enzymes involved in mitochondrial respiration (CitationLustbader et al 2004). ABAD can bind to the oligomeric Aβ42 present in the cortical mitochondria of APP transgenic mice (CitationYan et al 2007). Protease sensitivity assays suggest that Aβ gains access to the mitochondrial matrix rather than simply being adsorbed to the external surface of mitochondria (CitationCaspersen et al 2005). The interaction between Aβ and the mitochondria may explain how Aβ induces apoptosis and caspase activation (CitationIvins et al 1998; CitationWhite et al 2001; CitationLustbader et al 2004). Intracellullar Aβ may be derived from internalized extracellular Aβ or from intracellularly generated Aβ (CitationCasas et al 2004; CitationGomez-Ramos and Asuncion Moran 2007; CitationWegiel et al 2007). The presence of intracellular Aβ adds a further level of complexity to the mechanism of Aβ toxicity as this enables direct access to organelles that are vital for the function and viability of neurons. It is needless to say, that this has important implications for treatment strategies.

Conclusions

What can be expected in the forthcoming years? Some of the current therapeutic trials targeting Aβ may come to fruition (CitationGotz and Ittner 2008). With the advent of new tools it will likely become easier to discriminate Aβ conformations and hence allow establishing a defined role of specific conformers in toxicity (CitationHabicht et al 2007). The mode of Aβ uptake and/or binding by neurons and other cell types will be elucidated and interacting proteins, both under physiologic and pathologic conditions, will be identified. Finally, how Aβ and tau interact and contribute to disease will assist in the development of treatment strategies for AD and related disorders.

Acknowledgments

JG is a Medical Foundation Fellow and has been supported by the University of Sydney, the National Health and Medical Research Council (NHMRC), the Australian Research Council (ARC), the New South Wales Government through the Ministry for Science and Medical Research (BioFirst Program), the Nerve Research Foundation, the Medical Foundation (University of Sydney) and the Judith Jane Mason and Harold Stannett Williams Memorial Foundation.

Disclosures

The authors have no conflicts of interest to disclose.

References

  • AbramovAYCanevariLDuchenMR2004Calcium signals induced by amyloid beta peptide and their consequences in neurons and astrocytes in cultureBiochim Biophys Acta174281715590058
  • AmbroggioEEKimDHSeparovicF2005Surface behavior and lipid interaction of Alzheimer beta-amyloid peptide 1–42: a membrane-disrupting peptideBiophys J8827061315681641
  • AnandatheerthavaradaHKBiswasGRobinMA2003Mitochondrial targeting and a novel transmembrane arrest of Alzheimer’s amyloid precursor protein impairs mitochondrial function in neuronal cellsJ Cell Biol161415412695498
  • ArancioOZhangHPChenX2004RAGE potentiates Abeta-induced perturbation of neuronal function in transgenic miceEmbo J23409610515457210
  • ArendtT2003Synaptic plasticity and cell cycle activation in neurons are alternative effector pathways: the ‘Dr Jekyll and Mr. Hyde concept’ of Alzheimer’s disease or the yin and yang of neuroplasticityProg Neurobiol718324814687983
  • ArispeNRojasEPollardHB1993Alzheimer disease amyloid beta protein forms calcium channels in bilayer membranes: blockade by tromethamine and aluminumProc Natl Acad Sci U S A90567718380642
  • ArnoldSEHymanBTFloryJ1991The topographical and neuroanatomical distribution of neurofibrillary tangles and neuritic plaques in the cerebral cortex of patients with Alzheimer’s diseaseCereb Cortex1103161822725
  • BakerMMackenzieIRPickering-BrownSM2006Mutations in progranulin cause tau-negative frontotemporal dementia linked to chromosome 17Nature442916916862116
  • BarghornSNimmrichVStriebingerA2005Globular amyloid beta-peptide oligomer – a homogenous and stable neuropathological protein in Alzheimer’s diseaseJ Neurochem958344716135089
  • BhatiaRLinHLalR2000Fresh and globular amyloid beta protein (1–42) induces rapid cellular degeneration: evidence for AbetaP channel-mediated cellular toxicityFaseb J1412334310834945
  • BiXGallCMZhouJ2002Uptake and pathogenic effects of amyloid beta peptide 1–42 are enhanced by integrin antagonists and blocked by NMDA receptor antagonistsNeuroscience1128274012088742
  • BolmontTClavagueraFMeyer-LuehmannM2007Induction of tau pathology by intracerebral infusion of amyloid-beta -containing brain extract and by amyloid-beta deposition in APP x Tau transgenic miceAm J Pathol17120122018055549
  • BraakHBraakE1991Neuropathological stageing of Alzheimer-related changesActa Neuropathol (Berl)82239591759558
  • BraakHBraakE1995Staging of Alzheimer’s disease-related neurofibrillary changesNeurobiol Aging162718 discussion 8–847566337
  • BraithwaiteSPPaulSNairnAC2006Synaptic plasticity: one STEP at a timeTrends Neurosci29452816806510
  • CappaiRBarnhamKJ2008Delineating the mechanism of Alzheimer’s disease A beta peptide neurotoxicityNeurochem Res335263217762917
  • CasasCSergeantNItierJM2004Massive CA1/2 neuronal loss with intraneuronal and N-terminal truncated Abeta42 accumulation in a novel Alzheimer transgenic modelAm J Pathol165128930015466394
  • CaspersenCWangNYaoJ2005Mitochondrial Abeta: a potential focal point for neuronal metabolic dysfunction in Alzheimer’s diseaseFaseb J192040116210396
  • ChenFDavidDFerrariA2004aPosttranslational modifications of tau - Role in human tauopathies and modeling in transgenic animalsCurr Drug Targets55031515270197
  • ChenFWollmerMAHoerndliF2004bRole for glyoxalase I in Alzheimer’s diseaseProc Natl Acad Sci U S A10176879215128939
  • ChinJPalopJJYuGQ2004Fyn kinase modulates synaptotoxicity, but not aberrant sprouting, in human amyloid precursor protein transgenic miceJ Neurosci244692715140940
  • ChongYHShinYJLeeEO2006ERK1/2 activation mediates Abeta oligomer-induced neurotoxicity via caspase-3 activation and tau cleavage in rat organotypic hippocampal slice culturesJ Biol Chem281203152516714296
  • ClearyJPWalshDMHofmeisterJJ2005Natural oligomers of the amyloid-beta protein specifically disrupt cognitive functionNat Neurosci8798415608634
  • CoulsonEJ2006Does the p 75 neurotrophin receptor mediate Abeta-induced toxicity in Alzheimer’s disease?J Neurochem986546016893414
  • CrouchPJBlakeRDuceJA2005Copper-dependent inhibition of human cytochrome c oxidase by a dimeric conformer of amyloid-beta1–42J Neurosci25672915659604
  • CrutsMGijselinckIvan der ZeeJ2006Null mutations in progranulin cause ubiquitin-positive frontotemporal dementia linked to chromosome 17q21Nature442920416862115
  • DavidDHoerndliFGotzJ2005aFunctional Genomics meets neurodegenerative disorders Part I: Transcriptomic and proteomic technologyProg Neurobiol761536816168556
  • DavidDCHauptmannSScherpingI2005bProteomic and functional analysis reveal a mitochondrial dysfunction in P301L tau transgenic miceJ Biol Chem280238021415831501
  • DavidDCIttnerLMGehrigP2006β-Amyloid treatment of two complementary P301L tau-expressing Alzheimer’s disease models reveals similar deregulated cellular processesProteomics665667717111439
  • De FeliceFGVelascoPTLambertMP2007Abeta oligomers induce neuronal oxidative stress through an N-methyl-D-aspartate receptor-dependent mechanism that is blocked by the Alzheimer drug memantineJ Biol Chem2821159060117308309
  • DeaneRWuZSagareA2004LRP/amyloid beta-peptide interaction mediates differential brain efflux of Abeta isoformsNeuron433334415294142
  • DelacourteASergeantNChampainD2002Nonoverlapping but synergetic tau and APP pathologies in sporadic Alzheimer’s diseaseNeurology5939840712177374
  • DemuroAMinaEKayedR2005Calcium dysregulation and membrane disruption as a ubiquitous neurotoxic mechanism of soluble amyloid oligomersJ Biol Chem2801729430015722360
  • DetersNIttnerLMGotzJ2008Divergent phosphorylation pattern of tau in P301L tau transgenic miceEur J Neurosci281374718662339
  • DeviLPrabhuBMGalatiDF2006Accumulation of amyloid precursor protein in the mitochondrial import channels of human Alzheimer’s disease brain is associated with mitochondrial dysfunctionJ Neurosci2690576816943564
  • EckertADroseSBrandtU2008aOligomeric and fibrillar species of β-amyloid (Aβ42) both impair mitochondrial function in P301L tau transgenic miceJ Mol Med8612556718709343
  • EckertAHauptmannSScherpingI2008bSoluble beta-amyloid leads to mitochondrial defects in amyloid precursor protein and tau transgenic miceNeurodegener Dis5157918322377
  • EdbauerDWinklerERegulaJT2003Reconstitution of gamma-secretase activityNat Cell Biol5486812679784
  • FerrariAHoerndliFBaechiT2003Beta-amyloid induces PHF-like tau filaments in tissue cultureJ Biol Chem27840162812893817
  • GamesDAdamsDAlessandriniR1995Alzheimer-type neuropathology in transgenic mice overexpressing V717F beta-amyloid precursor protein [see comments]Nature37352377845465
  • GlennerGGWongCW1984Alzheimer’s disease: initial report of the purification and characterization of a novel cerebrovascular amyloid proteinBiochem Biophys Res Commun120885906375662
  • GoedertMWischikCMCrowtherRA1988Cloning and sequencing of the cDNA encoding a core protein of the paired helical filament of Alzheimer disease: identification as the microtubule-associated protein tauProc Natl Acad Sci U S A85405153131773
  • Gomez-RamosPAsuncion MoranM2007Ultrastructural localization of intraneuronal Abeta-peptide in Alzheimer disease brainsJ Alzheimers Dis1153917361035
  • GotzJ2001Tau and transgenic animal modelsBrain Res Brain Res Rev352668611423157
  • GotzJChenFBarmettlerR2001aTau filament formation in transgenic mice expressing P301L TauJ Biol Chem2765293411013246
  • GotzJChenFvan DorpeJ2001bFormation of neurofibrillary tangles in P301L tau transgenic mice induced by Abeta42 fibrilsScience2931491511520988
  • GotzJDetersNDoldissenA2007A decade of tau transgenic animal models and beyondBrain Pathol179110317493043
  • GotzJIttnerLM2008Animal models of Alzheimer’s disease and frontotemporal dementiaNat Rev Neurosci95324418568014
  • GotzJIttnerLMFandrichM2008Is tau aggregation toxic or protective: a sensible question in the absence of sensitive methods?J Alzheimers Dis14423918688093
  • GotzJIttnerLMKinsS2006Do axonal defects in tau and amyloid precursor protein transgenic animals model axonopathy in Alzheimer’s disease?J Neurochem98993100616787410
  • GotzJNitschRM2001Compartmentalized tau hyperphosphorylation and increased levels of kinases in transgenic miceNeuroreport1220071611435938
  • GotzJProbstASpillantiniMG1995Somatodendritic localization and hyperphosphorylation of tau protein in transgenic mice expressing the longest human brain tau isoformEmbo J141304137729409
  • GotzJSchildAHoerndliF2004aAmyloid-induced neurofibrillary tangle formation in Alzheimer’s disease: insight from transgenic mouse and tissue-culture modelsInt J Dev Neurosci224536515465275
  • GotzJStrefferJRDavidD2004bTransgenic animal models of Alzheimer’s disease and related disorders: Histopathology, behavior and therapyMol Psychiatry96648315052274
  • GotzJTolnayMBarmettlerR2001cOligodendroglial tau filament formation in transgenic mice expressing G272V tauEur J Neurosci1321314011422454
  • Graff-RadfordNRWoodruffBK2007Frontotemporal dementiaSemin Neurol27485717226741
  • HabichtGHauptCFriedrichRP2007Directed selection of a conformational antibody domain that prevents mature amyloid fibril formation by stabilizing Abeta protofibrilsProc Natl Acad Sci U S A10419232718042730
  • HardyJSelkoeDJ2002The amyloid hypothesis of Alzheimer’s disease: progress and problems on the road to therapeuticsScience297353612130773
  • HarperJDLansburyPTJr1997Models of amyloid seeding in Alzheimer’s disease and scrapie: mechanistic truths and physiological consequences of the time-dependent solubility of amyloid proteinsAnnu Rev Biochem663854079242912
  • HartleyDMWalshDMYeCP1999Protofibrillar intermediates of amyloid beta-protein induce acute electrophysiological changes and progressive neurotoxicity in cortical neuronsJ Neurosci1988768410516307
  • HoerndliFDavidDGotzJ2005Functional genomics meets neurodegenerative disorders. Part II: Application and data integrationProg Neurobiol761698816169146
  • HoerndliFJPelechSPapassotiropoulosA2007Abeta treatment and P301L tau expression in an Alzheimer’s disease tissue culture model act synergistically to promote aberrant cell cycle re-entryEur J Neurosci26607217587323
  • HoerndliFJToigoMSchildA2004Reference genes identified in SH-SY5Y cells using custom-made gene arrays with validation by quantitative polymerase chain reactionAnal Biochem335304115519568
  • HolscherC1998Possible causes of Alzheimer’s disease: amyloid fragments, free radicals, and calcium homeostasisNeurobiol Dis5129419848086
  • HortschanskyPSchroeckhVChristopeitT2005The aggregation kinetics of Alzheimer’s beta-amyloid peptide is controlled by stochastic nucleationProtein Sci141753915937275
  • HsiaoKChapmanPNilsenS1996Correlative memory deficits, Abeta elevation, and amyloid plaques in transgenic mice [see comments]Science274991028810256
  • HuttonMLendonCLRizzuP1998Association of missense and 5′-splice-site mutations in tau with the inherited dementia FTDP-17Nature39370259641683
  • IshiharaTHongMZhangB1999Age-dependent emergence and progression of a tauopathy in transgenic mice overexpressing the – human tau isoformNeuron247516210595524
  • IttnerLMFathTKeYD2008Parkinsonism and impaired axonal transport in a mouse model of frontotemporal dementiaProc Natl Acad Sci U S A105159971600218832465
  • IvinsKJBuiETCotmanCW1998Beta-amyloid induces local neurite degeneration in cultured hippocampal neurons: evidence for neuritic apoptosisNeurobiol Dis53657810069579
  • JanusCPearsonJMcLaurinJ2000A beta peptide immunization reduces behavioural impairment and plaques in a model of Alzheimer’s diseaseNature4089798211140685
  • JichaGABowserRKazamIG1997Alz-50 and MC-1, a new monoclonal antibody raised to paired helical filaments, recognize conformational epitopes on recombinant tauJ Neurosci Res48128329130141
  • KeilUBonertAMarquesCA2004Amyloid beta-induced changes in nitric oxide production and mitochondrial activity lead to apoptosisJ Biol Chem279503102015371443
  • KinsSCrameriAEvansDR2001Reduced PP2A activity induces hyperphosphorylation and altered compartmentalization of tau in transgenic miceJ Biol Chem2763819320011473109
  • KinsSKurosinskiPNitschRM2003Activation of the ERK and JNK signaling pathways caused by neuron specific inhibition of PP2A in transgenic miceAm J Pathol1638334312937125
  • KlyubinIWalshDMLemereCA2005Amyloid beta protein immunotherapy neutralizes Abeta oligomers that disrupt synaptic plasticity in vivoNat Med115566115834427
  • KulicLKurosinskiPChenF2006Active immunization trial in Abeta42-injected P301L tau transgenic miceNeurobiol Dis2250616289870
  • KurosinskiPGotzJ2002Glial cells under physiologic and pathological conditionsArch Neurol591524812374489
  • LambertMPBarlowAKChromyBA1998Diffusible, nonfibrillar ligands derived from Abeta-1 –42 are potent central nervous system neurotoxinsProc Natl Acad Sci U S A956448539600986
  • LauTLAmbroggioEETewDJ2006Amyloid-beta peptide disruption of lipid membranes and the effect of metal ionsJ Mol Biol3567597016403524
  • LeeGNewmanSTGardDL1998Tau interacts with src-family non-receptor tyrosine kinasesJ Cell Sci1113167779763511
  • LeeVMGoedertMTrojanowskiJQ2001Neurodegenerative tauopathiesAnnu Rev Neurosci2411215911520930
  • LesneSKohMTKotilinekL2006A specific amyloid-beta protein assembly in the brain impairs memoryNature440352716541076
  • LewisJDicksonDWLinW-L2001Enhanced neurofibrillary degeneration in transgenic mice expressing mutant Tau and APPScience29314879111520987
  • LimYAIttnerLMLimYL2008Human but not rat amylin shares neurotoxic properties with Abeta42 in long-term hippocampal and cortical culturesFEBS Lett58221889418486611
  • LinHBhatiaRLalR2001Amyloid beta protein forms ion channels: implications for Alzheimer’s disease pathophysiologyFaseb J1524334411689468
  • LustbaderJWCirilliMLinC2004ABAD directly links Abeta to mitochondrial toxicity in Alzheimer’s diseaseScience3044485215087549
  • ManczakMAnekondaTSHensonE2006Mitochondria are a direct site of A beta accumulation in Alzheimer’s disease neurons: implications for free radical generation and oxidative damage in disease progressionHum Mol Genet1514374916551656
  • MastersCLSimmsGWeinmanNA1985Amyloid plaque core protein in Alzheimer disease and Down syndromeProc Natl Acad Sci U S A82424593159021
  • MinogueAMSchmidAWFogartyMP2003Activation of the c-Jun N-terminal kinase signaling cascade mediates the effect of amyloid-beta on long term potentiation and cell death in hippocampus: a role for interleukin-1beta?J Biol Chem278279718012738769
  • MuckeLMasliahEYuGQ2000High-level neuronal expression of abeta 1–42 in wild-type human amyloid protein precursor transgenic mice: synaptotoxicity without plaque formationJ Neurosci204050810818140
  • NeculaMKayedRMiltonS2007Small molecule inhibitors of aggregation indicate that amyloid beta oligomerization and fibrillization pathways are independent and distinctJ Biol Chem282103112417284452
  • NeumannMMackenzieIRCairnsNJ2007TDP-43 in the ubiquitin pathology of frontotemporal dementia with VCP gene mutationsJ Neuropathol Exp Neurol66152717279000
  • NeumannMSampathuDMKwongLK2006Ubiquitinated TDP-43 in frontotemporal lobar degeneration and amyotrophic lateral sclerosisScience314130317023659
  • PatelVArayaRChatterjeeS2007Treatment and prevention of mental disorders in low-income and middle-income countriesLancet370991100517804058
  • PellarinRCaflischA2006Interpreting the aggregation kinetics of amyloid peptidesJ Mol Biol3608829216797587
  • PennanenLGotzJ2005Different tau epitopes define Abeta(42)-mediated tau insolubilityBiochem Biophys Res Commun337109710116226718
  • PennanenLWelzlHD‘AdamoP2004Accelerated extinction of conditioned taste aversion in P301L tau transgenic miceNeurobiol Dis15500915056457
  • PennanenLWolferDPNitschRM2006Impaired spatial reference memory and increased exploratory behavior in P301L tau transgenic miceGenes Brain Behav53697916879631
  • PeriniGDella-BiancaVPolitiV2002Role of p 75 neurotrophin receptor in the neurotoxicity by beta-amyloid peptides and synergistic effect of inflammatory cytokinesJ Exp Med1959071811927634
  • PikeCJBurdickDWalencewiczAJ1993Neurodegeneration induced by beta-amyloid peptides in vitro: the role of peptide assembly stateJ Neurosci131676878463843
  • PikeCJWalencewiczAJGlabeCG1991aAggregation-related toxicity of synthetic beta-amyloid protein in hippocampal culturesEur J Pharmacol20736781783006
  • PikeCJWalencewiczAJGlabeCG1991bIn vitro aging of beta-amyloid protein causes peptide aggregation and neurotoxicityBrain Res56331141786545
  • PollardHBArispeNRojasE1995Ion channel hypothesis for Alzheimer amyloid peptide neurotoxicityCell Mol Neurobiol15513268719038
  • PoorkajPBirdTDWijsmanE1998Tau is a candidate gene for chromosome 17 frontotemporal dementiaAnn Neurol43815259629852
  • ProbstAGotzJWiederholdKH2000Axonopathy and amyotrophy in mice transgenic for human four-repeat tau proteinActa Neuropathol (Berl)994698110805089
  • RapoportMDawsonHNBinderLI2002Tau is essential to beta-amyloid-induced neurotoxicityProc Natl Acad Sci U S A996364911959919
  • RobersonEDScearce-LevieKPalopJJ2007Reducing endogenous tau ameliorates amyloid beta-induced deficits in an Alzheimer’s disease mouse modelScience316750417478722
  • RocchiAPellegriniSSicilianoG2003Causative and susceptibility genes for Alzheimer’s disease: a reviewBrain Res Bull6112412788204
  • SelkoeDJ2002Alzheimer’s disease is a synaptic failureScience2987899112399581
  • ShibataMYamadaSKumarSR2000Clearance of Alzheimer’s amyloid-beta(1–40) peptide from brain by LDL receptor-related protein-1 at the blood-brain barrierJ Clin Invest10614899911120756
  • SmallDHMokSSBornsteinJC2001OPINION Alzheimer’s disease and Abeta toxicity: from top to bottomNat Rev Neurosci2595811484003
  • SmithDPCiccotostoGDTewDJ2007Concentration dependent Cu2+ induced aggregation and dityrosine formation of the Alzheimer’s disease amyloid-beta peptideBiochemistry4628819117297919
  • SnowdenJSBathgateDVarmaA2001Distinct behavioural profiles in frontotemporal dementia and semantic dementiaJ Neurol Neurosurg Psychiatry703233211181853
  • SnyderEMNongYAlmeidaCG2005Regulation of NMDA receptor trafficking by amyloid-betaNat Neurosci81051816025111
  • SotthibundhuASykesAMFoxB2008Beta-amyloid (1–42) induces neuronal death through the p 75 neurotrophin receptorJ Neurosci283941618400893
  • SpillantiniMGMurrellJRGoedertM1998Mutation in the tau gene in familial multiple system tauopathy with presenile dementiaProc Natl Acad Sci U S A957737419636220
  • SpittaelsKVan den HauteCVan DorpeJ1999Prominent axonopathy in the brain and spinal cord of transgenic mice overexpressing four-repeat human tau proteinAm J Pathol15521536510595944
  • StalderMPhinneyAProbstA1999Association of microglia with amyloid plaques in brains of APP23 transgenic miceAm J Pathol15416738410362792
  • StineWBJrDahlgrenKNKrafftGA2003In vitro characterization of conditions for amyloid-beta peptide oligomerization and fibrillogenesisJ Biol Chem278116122212499373
  • Sturchler-PierratCAbramowskiDDukeM1997Two amyloid precursor protein transgenic mouse models with Alzheimer disease-like pathologyProc Natl Acad Sci U S A9413287929371838
  • TakumaKYaoJHuangJ2005ABAD enhances Abeta-induced cell stress via mitochondrial dysfunctionFaseb J19597815665036
  • TrommerBLShahCYunSH2005ApoE isoform-specific effects on LTP: blockade by oligomeric amyloid-beta1–42Neurobiol Dis18758215649697
  • VassarRBennettBDBabu-KhanS1999Beta-secretase cleavage of Alzheimer’s amyloid precursor protein by the transmembrane aspartic protease BACEScience2867354110531052
  • VerdierYZarandiMPenkeB2004Amyloid beta-peptide interactions with neuronal and glial cell plasma membrane: binding sites and implications for Alzheimer’s diseaseJ Pept Sci102294815160835
  • WalshDMKlyubinIFadeevaJV2002Naturally secreted oligomers of amyloid beta protein potently inhibit hippocampal long-term potentiation in vivoNature416535911932745
  • WalshDMKlyubinIShankarGM2005The role of cell-derived oligomers of Abeta in Alzheimer’s disease and avenues for therapeutic interventionBiochem Soc Trans3310879016246051
  • WangQWalshDMRowanMJ2004Block of long-term potentiation by naturally secreted and synthetic amyloid beta-peptide in hippocampal slices is mediated via activation of the kinases c-Jun N-terminal kinase, cyclin-dependent kinase 5, and p 38 mitogen-activated protein kinase as well as metabotropic glutamate receptor type 5J Neurosci243370815056716
  • WederNDAzizRWilkinsK2007Frontotemporal dementias: a reviewAnn Gen Psychiatry61517565679
  • WegielJKuchnaINowickiK2007Intraneuronal Abeta immuno-reactivity is not a predictor of brain amyloidosis-beta or neurofibrillary degenerationActa Neuropathol11338940217237937
  • WhiteARGuirguisRBrazierMW2001Sublethal concentrations of prion peptide PrP106–126 or the amyloid beta peptide of Alzheimer’s disease activates expression of proapoptotic markers in primary cortical neuronsNeurobiol Dis829931611300725
  • WhiteJAManelliAMHolmbergKH2005Differential effects of oligomeric and fibrillar amyloid-beta 1–42 on astrocyte-mediated inflammationNeurobiol Dis184596515755672
  • YanSDChenXFuJ1996RAGE and amyloid-beta peptide neurotoxicity in Alzheimer’s diseaseNature382685918751438
  • YanYLiuYSorciM2007Surface plasmon resonance and nuclear magnetic resonance studies of ABAD-Abeta interactionBiochemistry4617243117253767
  • YanYWangC2007Abeta40 protects non-toxic Abeta42 monomer from aggregationJ Mol Biol3699091617481654
  • YanknerBADuffyLKKirschnerDA1990Neurotrophic and neurotoxic effects of amyloid beta protein: reversal by tachykinin neuropeptidesScience250279822218531
  • ZhangYHongYBounharY2003p 75 neurotrophin receptor protects primary cultures of human neurons against extracellular amyloid beta peptide cytotoxicityJ Neurosci2373859412917374

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.