Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2011 Mar 1.
Published in final edited form as: Expert Rev Neurother. 2010 May;10(5):683–691. doi: 10.1586/ern.10.27

Causes versus effects: the increasing complexities of Alzheimer’s disease pathogenesis

Siddhartha Mondragón-Rodríguez 1, Gustavo Basurto-Islas 1, Hyoung-gon Lee 2, George Perry 2,3, Xiongwei Zhu 2, Rudy J Castellani 4, Mark A Smith 2,
PMCID: PMC2922904  NIHMSID: NIHMS225674  PMID: 20420489

Abstract

Amyloid plaques and neurofibrillary tangles are the hallmarks of Alzheimer’s disease and have been the focus of disease etiology and pathogenesis. However, in the larger picture of a complex disease, the precise etiology of the lesions per se, as well as the clinical disease, remain to be defined. In this regard, to date no single process has been identified as a useful target and treatment efforts have shown no meaningful progress. Therefore, alternative ideas that may lead to new and effective treatment options are much needed.

Keywords: Alzheimer’s disease, amyloid, fibrillary deposits, oxidative stress, tau


Amyloid plaques and neurofibrillary tangles (NFTs) are pathological hallmarks of the progression of Alzheimer’s disease (AD). The questions of how, why and when they appear are, therefore, pertinent to understanding the disorder and, as a result, research efforts into the relationship between lesion and pathogenesis have been substantial. It is now known that specific molecular events are associated with the development of AD, although their exact relationship to lesion appearance is still controversial. Pathogenic mutations leading, for example, to amyloid-β (Aβ) deposition, are a primary piece of evidence [13], while a wide range of other metabolic abnormalities have been proposed [48]. During early AD stages, cell cycle deregulation has been related to the disease [9,10], which may also contribute to the protein aggregates. Other processes, such as oxidative damage to DNA and RNA, have also been implicated in the pathogenesis of AD [11], as well as lesion histogenesis [1215]. Still others suggest tau protein metabolism as the rate-limiting factor in disease, specifically phosphorylation, conformational changes and cleavage [1621]. A number of other mechanisms have been implicated, including such general mechanisms as apoptosis, inflammation and excitotoxicity. Nevertheless, while the disease effects are clear, the precise cause remains elusive, which is a concept that is sometimes forgotten in the enthusiasm over small molecules.

Genetic events: causative, but how?

Pathogenic mutations leading to familial autosomal-dominant AD involve the Aβ precursor protein (APP) gene [22,23], and presenilin 1 (PSEN1) and presenilin 2 (PSEN2) genes [24]. Apolipoprotein E (APOE) polymorphisms, sometimes referred to as familial late-onset AD, are not mutations per se, but are a significant predisposing factor [25,26]. In particular, the APOE gene has three isoforms (ε2, ε3 and ε4), with the ε4 isoform being the strongest predisposing allele [27]. APOE ε3/ε4 heterozygotes have two- to three-fold higher risk of developing AD compared with ε3/ε3 homozygotes, and ε4/ε4 homozygotes have more than twofold the risk of the ε3/ε4 genotype, while the presence of ε2 is somewhat protective [28,29]. Thus, three polymorphisms of the APOE regulatory region (−219 G/T, −427 C/T and −491 A/T) are implicated in AD. Notably, however, after adjustment of the APOE status, a large French case–control study (388 AD cases and 386 controls) differed from the hypothesis of an independent involvement of the APOE promoter region polymorphisms in AD risk [30]. However, recently, the combined gene effects between tau (intron 9, rs2471738) polymorphism (discussed later in this section) and the low-density lipoprotein receptor-related protein 1 (exon 3, rs1799986) polymorphism were found to be associated with a six-times higher risk of developing AD than subjects without these risk genotypes, suggesting that APOE polymorphisms could have a synergistic effect, but, more importantly, the APOE gene could also be promoting other AD factors, such as tau deposition [31]. Indeed, in an attempt to clarify the mechanism that relates APOE and AD, a transgenic model that carries the Arg-61 region, which is specific for domain interaction, a structural property of APOE4 that distinguishes it from APOE3 and predicted to contribute to the association of APOE4 with AD, has been developed. Interestingly, the mice display synaptic, functional and cognitive deficits, with ‘domain interaction’ postulated as the causative factor. Domain interaction further results in astrocytic dysfunction, which some view as an early event in APOE4-associated AD development. The authors suggest domain interaction as a possible therapeutic target for AD treatment [32].

In terms of strictly genetic factors, the autosomal dominant mutations in the PSEN1 and PSEN2 have also been widely implicated in the development of AD through the abnormal processing of Aβ protein, which leads to deposition as Aβ plaques [33,34]. Noteworthy also, is a PSEN1 double mutation (E318G plus G394V), which evidently affects γ-secretase activity [33], and a heterozygous missense mutation in exon 11 of the PSEN2 gene with a substitution from valine to methionine in position 393. However, experimentally, in vitro expression of PSEN2 V393M cDNA does not show a detectable increase in the secreted Aβ42/Aβ40 peptide ratio [35], suggesting a possible new mechanism for presenilin mutations apart from Aβ processing as traditionally proposed. Supporting this hypothesis, a PSEN1 polymorphism was found to be associated with decreased AD risk among carriers of the APOE4 allele [36]. Interestingly, presenilin mutations may also be useful for diagnosis, as cerebrospinal fluid analysis in sporadic AD cases showed a causative L424R PSEN1 mutation, as well as a correlation of increased total tau and phosphorylated-tau levels with decreased Aβ42 [37], suggesting again that presenilin mutations could be interacting with the other disease factors of AD, such as tau depositions. While a number of publications demonstrate that PSEN mutations result in the production of Aβ peptides that contribute to the neurodegeneration seen during AD, it is also clear that other mechanisms are involved.

Mutations in the Aβ sequence probably contribute to the formation of the pathological structures seen in AD, at least in the rare instances where they occur. The support for this concept includes recent data that showed single residue mutations P7S decreased the thermodynamic stability and greatly enhanced its capacity to form amyloid-like fibrils in systemic amyloid light-chain amyloidosis [38]. Similarly, another report shows that a single tyrosine residue difference in the stop codon in the gene encoding the prion protein C-terminus may significantly affect the site of Aβ deposits [39].

In addition to Aβ, deposits mainly composed of tau protein constitute the second prominent pathlogical feature of AD. Over 40 mutations in the tau gene have been identified, mainly in families with autosomal-dominant frontotemporal dementia and parkinsonism phenotype linked to chromosome 17, but also in sporadic families with Pick’s disease and AD. Specifically, mutations in tau have appeared to result in an imbalance in the ratio of the synthesized 3R and 4R tau isoforms, which stimulates protein aggregation and initiates neurodegeneration [40]. The aggregation of 4R-tau isoform has been found to be significantly higher than 3R isoform in in vitro aggregation assays [41] and is predominant in the typical deposits (NFTs) observed during AD [42].

Pursuing this further and incorporating many confounding risk factors for the disease, a genotyping of AD cases of single nucleotide polymorphisms has been published in a public database called Alzgene, naming close to 600 genes and 1900 polymorphisms [43]. The majority of the variations remain to be mechanistically related to the disease; nevertheless, this opens an interesting opportunity to charactize the relationships between genetics and AD development. Beyond individual genes as noted above, additional susceptibility regions, including those on chromosomes 9, 10 and 12, were found to be related to AD [44], although the precise risk remains unproven. Overall, since 1990, no more than six genes have been related to the development of AD; however, they only account for a minority of AD cases around the world. Notably, the majority of the genes discussed so far that are involved in the disease are related to the deposition of Aβ, lesions that characterize disease development and progression [45,46]. It should also be pointed out that although genetic factors do not account for all AD cases, their effects may be regarded as causal when they are present.

Fibrillar Aβ depositions: cause or effect?

Identification of the histochemical properties of senile plaque cores, with Aβ being the major protein component, understandably raised the possibility of a protein-mediated neurotoxic process that laid the foundation for a hypothesis that has changed only marginally in the past 20 years. The perceived strength in the amyloid cascade hypothesis is reflected in the scientific literature, which is voluminous and dominated by experimental studies that strictly adhere to the following assumptions: Aβ accumulates in senile plaques in the AD brain; specific point mutations in the gene for APP cause familial, early-onset AD; and, increased copy numbers of APP in some cases of Down’s syndrome lead to relatively early Aβ deposits and pathology generally associated with AD [47].

Hence, it is extrapolated that Aβ synthesis and deposition and, in particular, a relative increase in the synthesis and deposition of ‘pathogenic’ A β42, is the ‘rate-limiting’ factor in AD pathogenesis, while the accompanying pathologies (neurofibrillary pathology, neuronal loss and synaptic dysfunction) are secondary, end-organ phenomena. Such deposition is also closely related to genetic mutations in the presenilins, and, given that presenilins are necessary components of the γ-secretase complex in Notch proteolysis, thus provides evidence that presenilin-mediated β-secretase cleavage of APP (necessary along with β-secretase for cleavage of APP and production of A β42) is a causal process [4852]. That said, genetics might also be considered as only one aspect in the Aβ deposition cascade and, thus, its deposition is a compensatory response during the neurodegenerative scenario rather than a causal factor itself [46,53]. For example, by using cultured cortical neurons, Aβ induces apoptosis through the JNK–c-Jun–FasL–caspase-dependent extrinsic apoptotic pathway [54]. Furthermore, Aβ modulates redox factor-1, thus affecting both the cell death signaling pathways and DNA repair [55]. Aβ induces oxidative stress, predominantly via mitochondria, and also affects cholesterol balance [56], and NMDA and oligomeric Aβ42 can induce reactive oxygen species production in cortical neurons through activation of NADPH oxidase [57]. Aβ also binds with nanomolar affinity to the cellular prion protein, inducing synaptic dysfunction by affecting long-term potentiation [58]. The finding that Aβ oligomers decreased RACK1 distribution in the membrane fraction of cortical neurons suggests that the Aβ-induced loss of RACK1 distribution in the cell membrane may underlie the impairment of muscarinic regulation of PKC and GABAergic transmission [59]. Regarding this issue, application (1 h) of Aβ42 or Aβ25–35 (proto-)fibrils but not oligomers induced significant membrane depolarization of pyramidal cells and increased the activity of excitatory cell populations as measured by extracellular field recordings in the juvenile rodent brain, confirming the pathogenic significance of Aβ fibrils [60]. However, others suggest that monomers and oligomers are more toxic than the fibrillary structure itself, since upon Aβ42 fibrillization, neurotoxicity is reduced [61], suggesting that the aggregated form of Aβ is protective rather than pathological. Furthermore, oligomers directly extracted from AD patients are able to inhibit long-term potentiation, enhance long-term depression and reduce dendritic spine density in normal rodent hippocampus, suggesting that Aβ dimers are, as matter of fact, synaptotoxic [62]. It was also found that Aβ oligomers, by promoting reactive oxygen species, trigger neuronal damage through NMDA receptor-dependent calcium flux, emphasizing a synaptotoxic role [63]. Taken together, despite the dichotomy between the actions of the Aβ species, it is clear that Aβ is playing a crucial role during neurodegeneration, with actions that could be considered either a cause or effect.

Fibrillary tau deposition: cause or effect?

The NFTs, mainly comprised of the pathologically self-aggregated tau protein, were characterized around the same time [64] or even earlier than the Aβ deposits [65,66]; however, until the presence of NFTs along the hippocampal area was found to be congruous to the cognitive decline seen in AD patients, and even since, tau phosphorylation is generally regarded as a downstream event [6769]. Several post-translational events of the tau protein, such as phosphorylation [7072], conformational changes [73,74] and cleavage [18], are specifically noted in AD, possibly resulting from the following pathological behavior of modified tau protein: lack of microtubule associations, migrations to the cell cytoplasm and self-aggregation. A growing body of data supports the hypothesis of post-translational events as causative for the tau aggregation seen during AD. In this regard, phosphorylation of tau protein at residues 262, 293, 324 and 356 was found to modify the structural properties in repeats 1 and 2, in particular for Gln265–Lys26, which are in close proximity to Ser262 [75], the phosphorylation site that most strongly attenuates binding to microtubules [19]. Thus, phosphorylation of truncated tau protein at residues Ser396/404 was found to attenuate the ability of tau to stabilize the microtubules [76], suggesting that two or more events are needed for the protein to show its pathological effect. In the same vein, events that truncate tau protein have been found to affect the normal capacity of the tau protein to confer stability to the microtubules, and instead lead to self-aggregation into the NFTs [77,78]. In the same regard, the conformational non-native α-structure is related to the microtubule repeat tau protein domain and directly associates with filament development at the start of paired helical filaments formation [79].

While the relationship and pathological consequence between post-translational modification of tau protein and NFT formation is undeniable, what is causative is not so clear. In an attempt to address this issue, several studies suggested models accounting for the appearance of the three main events: phosphorylation, conformational changes and cleavage. Indeed, these translational events were proposed as sequential steps in a tau pathological processing model [8083]. These events appeared in a chronological fashion, thus fostering an interesting debate about which step was responsible for tau aggregation. Ultimately, phosphorylation was proposed as the earliest event in tau processing despite the need to explain its origins [83]. That is, phosphorylation itself requires other events, such as kinase deregulation, in which oxidative stress and metabolic homeostasis (discussed later) clearly play a role. In fact, some authors have suggested that the NFT deposits could just be a dynamic reservoir that reflects the kinetics of the disease [84].

From this perspective, it is clear that the modified tau protein is dependent on earlier events, and, therefore, its fibrillar deposits are not a cause but are rather an effect of the disease.

Oxidative stress & metabolic malfunction: is this the strongest candidate for a cause?

By definition, detection of damage resulting from reactive oxygen species is indicative of oxidative stress [85]. Reactive oxygen species are a byproduct of cellular oxidative metabolism and are generated in the mitochondria during oxidative phosphorylation as molecules with unpaired electrons, such as superoxide (O2). Indirect evidence of cellular oxidative stress is the increased expression of molecules involved in oxidant defense, such as heme oxygenase, superoxide dismutases, glutathione transferases and catalase [85]. It is important to note that neurons displaying signs of oxidative stress are not necessarily succumbing to oxidative stress, but may be adapting by way of oxidant defenses. Thus, neurodegenerative disorders where oxidative stress is postulated to play a role are associated with mechanisms that maintain a balance between oxidative stress and adaptation to this stress, ultimately reflecting the ability of living systems to dynamically regulate their defense mechanisms in response to oxidants. In the case of disease, the normal balance between production of and defense against oxidative stress has been challenged. These challenges become clear during the AD process, and include a strong presence of increased sulfhydryls, induction of heme oxygenase-1 and increased expression of Cu/Zn superoxide dismutase [8688], which indicates loss of homeostasis. Thus, such oxidative stress is reflected in several targets, such as DNA alteration [8991], kinase deregulation [9294], metabolic dysfunction [95,96] and cell cycle perturbation [97,98]. These phenomena provide a case for protein malfunction such as occurs with Aβ [99] and tau protein aggregation, and, therefore, it has been suggested that neurons actually respond to oxidative stress by increasing Aβ production [100], an increase associated with a consequent reduction in oxidative stress [101,102]. Similarly, Aβ itself can act as a genuine antioxidant, such as a potent superoxide dismutase [103]. Therefore, AD cases with APP mutations lose effective antioxidant capacity (due to mutation-driven protein dysfunction), while the extensive Aβ deposits themselves are signatures not of neurotoxicity per se, but of oxidative imbalance and an oxidative stress response. This is consistent with the data that Aβ deposits begin to appear around the age of 40 years [102].

Relating oxidative stress and tau protein, phosphorylation plays a pivotal role in redox balance, and it is, therefore, not surprising that oxidative stress, through activation of MAPK pathways, leads to phosphorylation [104106], or that conditions associated with chronic oxidant stress, such as AD, are associated with extensive phosphorylation of targets such as tau protein.

Notably, oxidative stress extends its effects beyond AD-related proteins; the formation of hydroxyl radicals affects all cellular macromolecules, including RNA [12]. While RNA alteration may lead to protein sequence anomalies [107], RNA destruction is more easily accommodated in cellular metabolism than is damage to DNA [13,108] or enzyme active site destruction.

Further relating oxidative stress to the cell cycle perturbation, it has been found that impairment of nucleolar function in response to stress is accompanied by perturbation of nucleolar structure, cell cycle arrest and stabilization of p53 [109]. The nucleolar target for downregulation of rDNA transcription is TIF-IA, an essential transcription factor that modulates the activity of RNA polymerase I. Upon stress, TIF-IA is phosphorylated by JNK2, and phosphorylation prevents TIF-IA from interaction with polymerase I, thereby impairing transcription complex formation and rRNA synthesis. Furthermore, stress-induced inactivation of TIF-IA is accompanied by translocation of TIF-IA from the nucleolus to the nucleoplasm [98]. Additionally, oxidative balance has been related to upregulation of Cdc42/Rac in select neuronal populations in cases of AD, which suggests an oncogenic process during AD [104,110,111].

Overall, the establishment that oxidative stress is linked to cell cycle re-entry is probably one of the earliest insults encountered in vulnerable neurons, as well as kinase deregulation, and with further research support, this may be considered a causal event [112,113].

Potential therapeutic targets

Cholinesterase inhibitors are one of the therapeutic approaches used for AD, for which the rationale is as follows: inhibition of cholinesterase increases extracellular levels of brain acetylcholine and improves cognitive processes as a consequence, significantly so in experimental animals [114]. However, when they are used in AD patients the results are much less efficient in terms of improvement; indeed, some researchers have tried to explain why this treatment does not show significant memory improvement, postulating that other events, such as alteration of other neurotransmitter systems and diffuse synaptic loss, are also playing an important role [115]. That said, attacking just one event would not be enough; therefore, some researchers have tested combined therapies, such as cholinergic and adrenergic drugs in rat models, but unfortunately, the data have not shown an advantage for their use as a therapeutic alternative [114].

Using Aβ as a pharmacological target has also been tested [116]. Many efforts have been made in order to approach the Aβ therapeutic target rationale. Indeed, data have emerged from novel strategies, such intracerebral sequestration of Aβ by an anti-Aβ monoclonal antibody 266 that inhibits the accumulation of multi-meric toxic Aβ. Such data, however, have only recently been tested in rat models [117]. In addition, the newly proposed DNA Aβ42 trimer immunization of mice that was developed to produce specific Th2-type antibodies [118], which in principle should reduce Aβ deposition, has not yet been demonstrated. On the other hand, it should be pointed out that many animal models have shown improvement after immunization, despite data in humans that have been disappointing [119].

Another specific drug that inhibits the Aβ aggregation is 3-amino-1-propane sulfonic acid (tramiprosate, 3-APS, Alzhemed); however, a recent study has shown that it also favors tau aggregation [120], reinforcing the notion that multi-targets should be carefully considered.

Antioxidant use has also entered into the field and, in this regard, population studies have reported that intake of antioxidants or poly-unsaturated fatty acids may be associated with a reduced incidence of dementia, and it has been reported that people who are mentally, socially and physically active have a reduced risk of cognitive impairment [121]. However, results from randomized trials of risk factor modification have been mixed and have not demonstrated any conclusive evidence of improvement for AD patients.

Since novel therapies have not yet proven successful, other strategies that address the increased complexities between all the events related to the disease should be pursued.

Summary – jumping to a risky but necessary conclusion

The original goal of this report was to discuss the various and probably related AD pathogenic theories, encompassing genetics, metabolic dysfunction, cell cycle re-entry, fibrillar deposits or oxidative stress, in an attempt to segregate those aspects that more accurately reflect cause verus those that are downstream effects. In spite of the numerous studies and the vast new knowledge of molecular pathogenesis, the empirical features of the pathological progression still appear most relevant and most accurately reflect the etiopathogenesis. In this regard, and according to the data resulting from the study of human disease, such as, cell models and transgenic models, all the events seems to be followed or even promoted by previous events, such that they are eliminated as potential causes of disease. Consistent with such an analysis, oxidative stress appears closest to a causal event, acknowledging that previous events may be promoting the oxidative stress itself. However, it should be noted that while not all the events are, strictly speaking, causal, they may nevertheless affect disease progression (Figure 1). This may explain why the efforts in treatment have been singularly unsuccessful, but nevertheless may be useful in a ‘cocktail’ approach [122]. Ultimately, we propose that future attempts at AD treatment should recognize the multitude of complex, yet inter-related pathogenic mechanisms, and perhaps new promising data will emerge, if not by hypothesis-driven constructs, then by accident.

Figure 1. Tentative model for protein deposition and oxidative stress relationship.

Figure 1

Soluble amyloid-β oligomers could activate ionic channels (Ca2+) leading to an increase of Ca2+ entry prior to amyloid-β fibrilization. Increased Ca2+ additionally influences mitochondria and metabolic enzyme production of reactive oxygen species that in turn could affect proteins, DNA and lipids, and subsequently lead to necrosis, apoptosis and altered proteolysis – all events directly relating to tau processing and deposition. Additionally, reactive oxygen species abnormally activate kinases and phosphatases, which also directly affect tau processing and deposition. AA: Arachidonic acid; JNK: c-Jun N-terminal kinase; NOS: Nitric oxide synthase; PLA2: Phospholipase A2; PP2A: Protein phosphatase 2A; SOD: Superoxide dismutase.

Expert commentary

For more than a decade, the field of AD has focused on two hypotheses positing Aβ and/or tau phosphorylation as key pathogenetic mediators. While hypothetical constructs, these hypotheses have dictated the direction of the field. Recent experimental data, coupled with a number of clinical trial failures targeting these hypotheses, are now casting doubt that these hypotheses are fundamentally flawed. That is, the role of upstream mediators of disease, such as oxidative stress, should be investigated with the hope of establishing an effective means to reduce and/or prevent disease.

Five-year view

As much as one swallow does not make a summer, one failure should not lead to an abandonment of amyloid as a target. However, over the next 5 years, the amyloid hypothesis will have been repeatedly tested in clinical trials. Multiple failures will not be for a lack of effort, and should therefore lead investigators to re-organize their mindset.

Additionally, over the next 5 years, it is likely that clinical trial data from other therapeutic targets will be available. Success in any one of these areas may lead to new directions and increased attention to ‘new’ and perhaps more promising therapeutic pathways.

Key issues.

  • Distinguishing causal phenomena in Alzheimer’s disease (AD) from their associated epiphenomena is an essential step toward establishing a universal avenue to disease control.

  • The pathological hallmarks that characterize the disease, such as Aβ and neurofibrillary/tau tangles, are not primary events in disease pathogenesis, although they indeed mediate neurodegeneration.

  • Although only relevant in cases of familial, early onset AD, genetic mutations in APP, PSEN1/2 and APOE genes are responsible for the origins of disease pathogenesis (i.e., oxidative stress, cell cycle aberrations and amyloid-β [Aβ] deposition).

  • Aβ is initially instituted in the brain as a compensatory response to neuronal environmental changes (unless otherwise produced from point mutations), and it induces its destructive effects only after substantial aggregation and overaccumulation.

  • Aβ is thus a secondary aspect to AD development and not a primary event.

  • Similarly, tau phosphorylation and deposition is merely a secondary event in disease pathogenesis that results from environmental changes in the AD brain.

  • Oxidative stress is the best candidate for a primary causal factor in the neurodegeneration in AD.

  • Imbalances in the generation of reactive oxidative species, due to high demand of the brain for oxidative phosphorylation, induce oxidative stress on neuronal macromolecules (i.e., DNA, RNA and proteins, phospholipids) and require the brain to initiate antioxidant compensatory mechanisms that render the brain vulnerable to other damage.

  • This damage includes cell cycle aberrations, Aβ overaccumulation and neurofibrillary tangle aggregation: events that then mediate further neurodegeneration via autoimmune response, neuroinflammation, further oxidative stress and so on.

  • Oxidative stress and its immediate products are thus an avenue for future therapeutic intervention.

Footnotes

For reprint orders, please contact reprints@expert-reviews.com

Financial & competing interests disclosure

This work was funded by NIH grant AG026151. Mark A Smith is, or has in the past been, a paid consultant for, owns equity or stock options in and/or receives grant funding from Anavex, Canopus BioPharma, Medivation, Neurotez, Neuropharm, Panacea Pharmaceuticals and Voyager Pharmaceuticals. George Perry is, or has in the past been, a paid consultant for and/or owns equity or stock options in Neurotez Pharmaceuticals, Panacea Pharmaceuticals, Takeda Pharmaceuticals and Voyager Pharmaceuticals. Xiongwei Zhu is a paid Advisory Board member for Medivation. The authors have no other relevant affiliations or financial involvement with any organization or entity with a financial interest in or financial conflict with the subject matter or materials discussed in the manuscript apart from those disclosed.

No writing assistance was utilized in the production of this manuscript.

References

  • 1.Zekanowski C, Styczynska M, Peplonska B, et al. Mutations in presenilin 1, presenilin 2 and amyloid precursor protein genes in patients with early-onset Alzheimer’s disease in Poland. Exp Neurol. 2003;184(2):991–996. doi: 10.1016/S0014-4886(03)00384-4. [DOI] [PubMed] [Google Scholar]
  • 2.Zekanowski C, Religa D, Graff C, Filipek S, Kuznicki J. Genetic aspects of Alzheimer’s disease. Acta Neurobiol Exp (Wars) 2004;64(1):19–31. doi: 10.55782/ane-2004-1488. [DOI] [PubMed] [Google Scholar]
  • 3.Zawia NH, Lahiri DK, Cardozo-Pelaez F. Epigenetics, oxidative stress, and Alzheimer disease. Free Radic Biol Med. 2009;46(9):1241–1249. doi: 10.1016/j.freeradbiomed.2009.02.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Smith MA, Nunomura A, Zhu X, Takeda A, Perry G. Metabolic, metallic, and mitotic sources of oxidative stress in Alzheimer disease. Antioxid Redox Signal. 2000;2(3):413–420. doi: 10.1089/15230860050192198. [DOI] [PubMed] [Google Scholar]
  • 5.Rapoport SI, Horwitz B, Grady CL, Haxby JV, DeCarli C, Schapiro MB. Abnormal brain glucose metabolism in Alzheimer’s disease, as measured by position emission tomography. Adv Exp Med Biol. 1991;29(1):231–248. doi: 10.1007/978-1-4684-5931-9_18. [DOI] [PubMed] [Google Scholar]
  • 6.Matsuda H. Cerebral blood flow and metabolic abnormalities in Alzheimer’s disease. Ann Nucl Med. 2001;15(2):85–92. doi: 10.1007/BF02988596. [DOI] [PubMed] [Google Scholar]
  • 7.Blass JP. Cerebrometabolic abnormalities in Alzheimer’s disease. Neurol Res. 2003;25(6):556–566. doi: 10.1179/016164103101201995. [DOI] [PubMed] [Google Scholar]
  • 8.Ackl N, Ising M, Schreiber YA, Atiya M, Sonntag A, Auer DP. Hippocampal metabolic abnormalities in mild cognitive impairment and Alzheimer’s disease. Neurosci Lett. 2005;384(1–2):23–28. doi: 10.1016/j.neulet.2005.04.035. [DOI] [PubMed] [Google Scholar]
  • 9.Webber KM, Raina AK, Marlatt MW, et al. The cell cycle in Alzheimer disease: a unique target for neuropharmacology. Mech Ageing Dev. 2005;126(10):1019–1025. doi: 10.1016/j.mad.2005.03.024. [DOI] [PubMed] [Google Scholar]
  • 10.Woods J, Snape M, Smith MA. The cell cycle hypothesis of Alzheimer’s disease: suggestions for drug development. Biochim Biophys Acta. 2007;1772(4):503–508. doi: 10.1016/j.bbadis.2006.12.004. [DOI] [PubMed] [Google Scholar]
  • 11.Lee KW, Kim JB, Seo JS, et al. Behavioral stress accelerates plaque pathogenesis in the brain of Tg2576 mice via generation of metabolic oxidative stress. J Neurochem. 2009;108(1):165–175. doi: 10.1111/j.1471-4159.2008.05769.x. [DOI] [PubMed] [Google Scholar]
  • 12.Ma YS, Wu SB, Lee WY, Cheng JS, Wei YH. Response to the increase of oxidative stress and mutation of mitochondrial DNA in aging. Biochim Biophys Acta. 2009;1790(10):1021–1029. doi: 10.1016/j.bbagen.2009.04.012. [DOI] [PubMed] [Google Scholar]
  • 13.Gabbita SP, Lovell MA, Markesbery WR. Increased nuclear DNA oxidation in the brain in Alzheimer’s disease. J Neurochem. 1998;71(5):2034–2040. doi: 10.1046/j.1471-4159.1998.71052034.x. [DOI] [PubMed] [Google Scholar]
  • 14.Kadioglu E, Sardas S, Aslan S, Isik E, Esat Karakaya A. Detection of oxidative DNA damage in lymphocytes of patients with Alzheimer’s disease. Biomarkers. 2004;9(2):203–209. doi: 10.1080/13547500410001728390. [DOI] [PubMed] [Google Scholar]
  • 15.Lovell MA, Markesbery WR. Oxidatively modified RNA in mild cognitive impairment. Neurobiol Dis. 2008;29(2):169–175. doi: 10.1016/j.nbd.2007.07.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Kovacech B, Zilka N, Novak M. New age of neuroproteomics in Alzheimer’s disease research. Cell Mol Neurobiol. 2009;29(6–7):799–805. doi: 10.1007/s10571-009-9358-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Ghoshal N, Garcia-Sierra F, Wuu J, et al. Tau conformational changes correspond to impairments of episodic memory in mild cognitive impairment and Alzheimer’s disease. Exp Neurol. 2002;177(2):475–493. doi: 10.1006/exnr.2002.8014. [DOI] [PubMed] [Google Scholar]
  • 18.Gamblin TC, Chen F, Zambrano A, et al. Caspase cleavage of tau: linking amyloid and neurofibrillary tangles in Alzheimer’s disease. Proc Natl Acad Sci USA. 2003;100(17):10032–10037. doi: 10.1073/pnas.1630428100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Biernat J, Gustke N, Drewes G, Mandelkow EM, Mandelkow E. Phosphorylation of Ser262 strongly reduces binding of tau to microtubules, distinction between PHF-like immunoreactivity and microtubule binding. Neuron. 1993;11(1):153–163. doi: 10.1016/0896-6273(93)90279-z. [DOI] [PubMed] [Google Scholar]
  • 20.Alonso A, Zaidi T, Novak M, Grundke-Iqbal I, Iqbal K. Hyperphosphorylation induces self-assembly of tau into tangles of paired helical filaments/straight filaments. Proc Natl Acad Sci USA. 2001;98(12):6923–6928. doi: 10.1073/pnas.121119298. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Avila J. Tau phosphorylation and aggregation in Alzheimer’s disease pathology. FEBS Lett. 2006;580(12):2922–2927. doi: 10.1016/j.febslet.2006.02.067. [DOI] [PubMed] [Google Scholar]
  • 22.Di Fede G, Catania M, Morbin M, et al. A recessive mutation in the APP gene with dominant-negative effect on amyloidogenesis. Science. 2009;323(5920):1473–1477. doi: 10.1126/science.1168979. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Bernardi L, Geracitano S, Colao R, et al. AβPP A713T mutation in late onset Alzheimer’s disease with cerebrovascular lesions. J Alzheimers Dis. 2009;17(2):383–389. doi: 10.3233/JAD-2009-1061. [DOI] [PubMed] [Google Scholar]
  • 24.Larner AJ, Doran M. Genotype–phenotype relationships of presenilin-1 mutations in Alzheimer’s disease: an update. J Alzheimers Dis. 2009;17(2):259–265. doi: 10.3233/JAD-2009-1042. [DOI] [PubMed] [Google Scholar]
  • 25.Aleshkov S, Abraham CR, Zannis VI. Interaction of nascent ApoE2, ApoE3, and ApoE4 isoforms expressed in mammalian cells with amyloid peptide β1–40. Relevance to Alzheimer’s disease. Biochemistry. 1997;36(34):10571–10580. doi: 10.1021/bi9626362. [DOI] [PubMed] [Google Scholar]
  • 26.Adalbert R, Gilley J, Coleman MP. Aβ, tau and ApoE4 in Alzheimer’s disease, the axonal connection. Trends Mol Med. 2007;13(4):135–142. doi: 10.1016/j.molmed.2007.02.004. [DOI] [PubMed] [Google Scholar]
  • 27.Bu G. Apolipoprotein E and its receptors in Alzheimer’s disease: pathways, pathogenesis and therapy. Nat Rev Neurosci. 2009;10(5):333–344. doi: 10.1038/nrn2620. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Aggarwal NT, Wilson RS, Beck TL, Bienias JL, Berry-Kravis E, Bennett DA. The apolipoprotein E ε4 allele and incident Alzheimer’s disease in persons with mild cognitive impairment. Neurocase. 2005;11(1):3–7. doi: 10.1080/13554790490903038. [DOI] [PubMed] [Google Scholar]
  • 29.Adroer R, Santacruz P, Blesa R, Lopez-Pousa S, Ascaso C, Oliva R. Apolipoprotein E4 allele frequency in Spanish Alzheimer and control cases. Neurosci Lett. 1995;189(3):182–186. doi: 10.1016/0304-3940(95)11480-k. [DOI] [PubMed] [Google Scholar]
  • 30.Zurutuza L, Verpillat P, Raux G, et al. APOE promoter polymorphisms do not confer independent risk for Alzheimer‘s disease in a French population. Eur J Hum Genet. 2000;8(9):713–716. doi: 10.1038/sj.ejhg.5200513. [DOI] [PubMed] [Google Scholar]
  • 31.Vazquez-Higuera JL, Mateo I, Sanchez-Juan P, et al. Genetic interaction between tau and the apolipoprotein E receptor LRP1 increases Alzheimer’s disease risk. Dement Geriatr Cogn Disord. 2009;28(2):116–120. doi: 10.1159/000234913. [DOI] [PubMed] [Google Scholar]
  • 32.Zhong N, Weisgraber KH. Understanding the basis for the association of apoE4 with Alzheimer’s disease: opening the door for therapeutic approaches. Curr Alzheimer Res. 2009;6(5):415–418. doi: 10.2174/156720509789207921. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Batelli S, Albani D, Prato F, et al. Early-onset Alzheimer disease in an Italian family with presenilin-1 double mutation E318G and G394V. Alzheimer Dis Assoc Disord. 2008;22(2):184–187. doi: 10.1097/WAD.0b013e31815a9dec. [DOI] [PubMed] [Google Scholar]
  • 34.Brouwers N, Sleegers K, Van Broeckhoven C. Molecular genetics of Alzheimer’s disease: an update. Ann Med. 2008;40(8):562–583. doi: 10.1080/07853890802186905. [DOI] [PubMed] [Google Scholar]
  • 35.Lindquist SG, Hasholt L, Bahl JM, et al. A novel presenilin 2 mutation (V393M) in early-onset dementia with profound language impairment. Eur J Neurol. 2008;15(10):1135–1139. doi: 10.1111/j.1468-1331.2008.02256.x. [DOI] [PubMed] [Google Scholar]
  • 36.Martinez-Garcia A, Aldudo J, Recuero M, et al. Presenilin 1 polymorphism associated with Alzheimer’s disease in apolipoprotein E4 carriers. Dement Geriatr Cogn Disord. 2008;26(5):440–444. doi: 10.1159/000165685. [DOI] [PubMed] [Google Scholar]
  • 37.de Bot ST, Kremer HP, Dooijes D, Verbeek MM. CSF studies facilitate DNA diagnosis in familial Alzheimer’s disease due to a presenilin-1 mutation. J Alzheimers Dis. 2009;17(1):53–57. doi: 10.3233/JAD-2009-1038. [DOI] [PubMed] [Google Scholar]
  • 38.Hernandez-Santoyo A, Del Pozo Yauner L, Fuentes-Silva D, et al. A single mutation at the sheet switch region results in conformational changes favoring λ6 light-chain fibrillogenesis. J Mol Biol. 2010;396(2):280–292. doi: 10.1016/j.jmb.2009.11.038. [DOI] [PubMed] [Google Scholar]
  • 39.Jansen C, Parchi P, Capellari S, et al. Prion protein amyloidosis with divergent phenotype associated with two novel nonsense mutations in PRNP. Acta Neuropathol. 2009;119(2):189–197. doi: 10.1007/s00401-009-0609-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Kowalska A. The genetics of dementias. Part 1: molecular basis of frontotemporal dementia and parkinsonism linked to chromosome 17 (FTDP-17) Postepy Hig Med Dosw (Online) 2009;63:278–286. [PubMed] [Google Scholar]
  • 41.Sadik G, Tanaka T, Kato K, Yanagi K, Kudo T, Takeda M. Differential interaction and aggregation of 3-repeat and 4-repeat tau isoforms with 14-13-3ζ protein. Biochem Biophys Res Commun. 2009;383(1):37–41. doi: 10.1016/j.bbrc.2009.03.107. [DOI] [PubMed] [Google Scholar]
  • 42.Espinoza M, de Silva R, Dickson DW, Davies P. Differential incorporation of tau isoforms in Alzheimer’s disease. J Alzheimers Dis. 2008;14(1):1–16. doi: 10.3233/jad-2008-14101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Avramopoulos D. Genetics of Alzheimer’s disease: recent advances. Genome Med. 2009;1(3):34. doi: 10.1186/gm34. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Liang X, Schnetz-Boutaud N, Kenealy SJ, et al. Covariate analysis of late-onset Alzheimer disease refines the chromosome 12 locus. Mol Psychiatry. 2006;11(3):280–285. doi: 10.1038/sj.mp.4001766. [DOI] [PubMed] [Google Scholar]
  • 45.Lee HG, Casadesus G, Zhu X, Joseph JA, Perry G, Smith MA. Perspectives on the amyloid-β cascade hypothesis. J Alzheimers Dis. 2004;6(2):137–145. doi: 10.3233/jad-2004-6205. [DOI] [PubMed] [Google Scholar]
  • 46.Lee HG, Castellani RJ, Zhu X, Perry G, Smith MA. Amyloid-β in Alzheimer’s disease: the horse or the cart? Pathogenic or protective? Int J Exp Pathol. 2005;86(3):133–138. doi: 10.1111/j.0959-9673.2005.00429.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Hardy J, Selkoe DJ. The amyloid hypothesis of Alzheimer’s disease: progress and problems on the road to therapeutics. Science. 2002;297(5580):353–356. doi: 10.1126/science.1072994. [DOI] [PubMed] [Google Scholar]
  • 48.Sato N, Imaizumi K, Manabe T, et al. Increased production of β-amyloid and vulnerability to endoplasmic reticulum stress by an aberrant spliced form of presenilin 2. J Biol Chem. 2001;276(3):2108–2114. doi: 10.1074/jbc.M006886200. [DOI] [PubMed] [Google Scholar]
  • 49.Wolfe MS, Xia W, Ostaszewski BL, Diehl TS, Kimberly WT, Selkoe DJ. Two transmembrane aspartates in presenilin-1 required for presenilin endoproteolysis and γ-secretase activity. Nature. 1999;398(6727):513–517. doi: 10.1038/19077. [DOI] [PubMed] [Google Scholar]
  • 50.Struhl G, Greenwald I. Presenilin is required for activity and nuclear access of Notch in Drosophila. Nature. 1999;398(6727):522–525. doi: 10.1038/19091. [DOI] [PubMed] [Google Scholar]
  • 51.Ye Y, Lukinova N, Fortini ME. Neurogenic phenotypes and altered Notch processing in Drosophila presenilin mutants. Nature. 1999;398(6727):525–529. doi: 10.1038/19096. [DOI] [PubMed] [Google Scholar]
  • 52.De Strooper B, Annaert W, Cupers P, et al. A presenilin-1-dependent γ-secretase-like protease mediates release of Notch intracellular domain. Nature. 1999;398(6727):518–522. doi: 10.1038/19083. [DOI] [PubMed] [Google Scholar]
  • 53.Lee HG, Zhu X, Castellani RJ, Nunomura A, Perry G, Smith MA. Amyloid-β in Alzheimer disease: the null versus the alternate hypotheses. J Pharmacol Exp Ther. 2007;321(3):823–829. doi: 10.1124/jpet.106.114009. [DOI] [PubMed] [Google Scholar]
  • 54.Li LM, Liu QH, Qiao JT, Zhang C. Aβ31–35-induced neuronal apoptosis is mediated by JNK-dependent extrinsic apoptosis pathway. Neurosci Bull. 2009;25(6):361–366. doi: 10.1007/s12264-009-0629-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Tan Z, Shi L, Schreiber SS. Differential expression of redox factor-1 associated with β-amyloid-mediated neurotoxicity. Open Neurosci J. 2009;3:26–34. doi: 10.2174/1874082000903010026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Colell A, Fernandez A, Fernandez-Checa JC. Mitochondria, cholesterol and amyloid β peptide: a dangerous trio in Alzheimer disease. J Bioenerg Biomembr. 2009;41(5):417–423. doi: 10.1007/s10863-009-9242-6. [DOI] [PubMed] [Google Scholar]
  • 57.Shelat PB, Chalimoniuk M, Wang JH, et al. Amyloid β peptide and NMDA induce ROS from NADPH oxidase and AA release from cytosolic phospholipase A2 in cortical neurons. J Neurochem. 2008;106(1):45–55. doi: 10.1111/j.1471-4159.2008.05347.x. [DOI] [PubMed] [Google Scholar]
  • 58.Lauren J, Gimbel DA, Nygaard HB, Gilbert JW, Strittmatter SM. Cellular prion protein mediates impairment of synaptic plasticity by amyloid-β oligomers. Nature. 2009;457(7233):1128–1132. doi: 10.1038/nature07761. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Liu W, Dou F, Feng J, Yan Z. RACK1 is involved in β-amyloid impairment of muscarinic regulation of GABAergic transmission. Neurobiol Aging. 2009 doi: 10.1016/j.neurobiolaging. 2009.10.017. (Epub ahead of print) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Minkeviciene R, Rheims S, Dobszay MB, et al. Amyloid β-induced neuronal hyperexcitability triggers progressive epilepsy. J Neurosci. 2009;29(11):3453–3462. doi: 10.1523/JNEUROSCI.5215-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Feng Y, Yang SG, Du XT, et al. Ellagic acid promotes Aβ42 fibrillization and inhibits Aβ42-induced neurotoxicity. Biochem Biophys Res Commun. 2009;390(4):1250–1254. doi: 10.1016/j.bbrc.2009.10.130. [DOI] [PubMed] [Google Scholar]
  • 62.Shankar GM, Li S, Mehta TH, et al. Amyloid-β protein dimers isolated directly from Alzheimer’s brains impair synaptic plasticity and memory. Nat Med. 2008;14(8):837–842. doi: 10.1038/nm1782. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.De Felice FG, Velasco PT, Lambert MP, et al. Aβ oligomers induce neuronal oxidative stress through an N-methyl-D-aspartate receptor-dependent mechanism that is blocked by the Alzheimer drug memantine. J Biol Chem. 2007;282(15):11590–11601. doi: 10.1074/jbc.M607483200. [DOI] [PubMed] [Google Scholar]
  • 64.Wood JG, Mirra SS, Pollock NJ, Binder LI. Neurofibrillary tangles of Alzheimer disease share antigenic determinants with the axonal microtubule-associated protein tau (tau) Proc Natl Acad Sci USA. 1986;83(11):4040–4043. doi: 10.1073/pnas.83.11.4040. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Zhan SS, Veerhuis R, Kamphorst W, Eikelenboom P. Distribution of β amyloid associated proteins in plaques in Alzheimer’s disease and in the non-demented elderly. Neurodegeneration. 1995;4(3):291–297. doi: 10.1016/1055-8330(95)90018-7. [DOI] [PubMed] [Google Scholar]
  • 66.Allsop D, Wong CW, Ikeda S, Landon M, Kidd M, Glenner GG. Immunohistochemical evidence for the derivation of a peptide ligand from the amyloid β-protein precursor of Alzheimer disease. Proc Natl Acad Sci USA. 1988;85(8):2790–2794. doi: 10.1073/pnas.85.8.2790. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Braak H, Braak E. Demonstration of amyloid deposits and neurofibrillary changes in whole brain sections. Brain Pathol. 1991;1(3):213–216. doi: 10.1111/j.1750-3639.1991.tb00661.x. [DOI] [PubMed] [Google Scholar]
  • 68.Braak H, Braak E. Neuropathological stageing of Alzheimer-related changes. Acta Neuropathol. 1991;82(4):239–259. doi: 10.1007/BF00308809. [DOI] [PubMed] [Google Scholar]
  • 69.Braak H, Braak E. Morphological criteria for the recognition of Alzheimer’s disease and the distribution pattern of cortical changes related to this disorder. Neurobiol Aging. 1994;15(3):355–356. doi: 10.1016/0197-4580(94)90032-9. discussion 379–380. [DOI] [PubMed] [Google Scholar]
  • 70.Alonso AC, Grundke-Iqbal I, Iqbal K. Alzheimer’s disease hyperphosphorylated tau sequesters normal tau into tangles of filaments and disassembles microtubules. Nat Med. 1996;2(7):783–787. doi: 10.1038/nm0796-783. [DOI] [PubMed] [Google Scholar]
  • 71.Alonso AC, Zaidi T, Grundke-Iqbal I, Iqbal K. Role of abnormally phosphorylated tau in the breakdown of microtubules in Alzheimer disease. Proc Natl Acad Sci USA. 1994;91(12):5562–5566. doi: 10.1073/pnas.91.12.5562. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Augustinack JC, Schneider A, Mandelkow EM, Hyman BT. Specific tau phosphorylation sites correlate with severity of neuronal cytopathology in Alzheimer’s disease. Acta Neuropathol. 2002;103(1):26–35. doi: 10.1007/s004010100423. [DOI] [PubMed] [Google Scholar]
  • 73.Hyman BT, Van Hoesen GW, Wolozin BL, Davies P, Kromer LJ, Damasio AR. Alz-50 antibody recognizes Alzheimer-related neuronal changes. Ann Neurol. 1988;23(4):371–379. doi: 10.1002/ana.410230410. [DOI] [PubMed] [Google Scholar]
  • 74.Jicha GA, Bowser R, Kazam IG, Davies P. Alz-50 and MC-1, a new monoclonal antibody raised to paired helical filaments, recognize conformational epitopes on recombinant tau. J Neurosci Res. 1997;48(2):128–132. doi: 10.1002/(sici)1097-4547(19970415)48:2<128::aid-jnr5>3.0.co;2-e. [DOI] [PubMed] [Google Scholar]
  • 75.Fischer D, Mukrasch MD, Biernat J, et al. Conformational changes specific for pseudophosphorylation at serine 262 selectively impair binding of tau to microtubules. Biochemistry. 2009;48(42):10047–10055. doi: 10.1021/bi901090m. [DOI] [PubMed] [Google Scholar]
  • 76.Ding H, Matthews TA, Johnson GV. Site-specific phosphorylation and caspase cleavage differentially impact tau-microtubule interactions and tau aggregation. J Biol Chem. 2006;281(28):19107–19114. doi: 10.1074/jbc.M511697200. [DOI] [PubMed] [Google Scholar]
  • 77.Zilka N, Filipcik P, Koson P, et al. Truncated tau from sporadic Alzheimer’s disease suffices to drive neurofibrillary degeneration in vivo. FEBS Lett. 2006;580(15):3582–3588. doi: 10.1016/j.febslet.2006.05.029. [DOI] [PubMed] [Google Scholar]
  • 78.Iqbal K, Novak M. From tangles to tau protein. Bratisl Lek Listy. 2006;107(9–10):341–342. [PubMed] [Google Scholar]
  • 79.Hiraoka S, Yao TM, Minoura K, et al. Conformational transition state is responsible for assembly of microtubule-binding domain of tau protein. Biochem Biophys Res Commun. 2004;315(3):659–663. doi: 10.1016/j.bbrc.2004.01.107. [DOI] [PubMed] [Google Scholar]
  • 80.Guillozet-Bongaarts AL, Garcia-Sierra F, Reynolds MR, et al. Tau truncation during neurofibrillary tangle evolution in Alzheimer’s disease. Neurobiol Aging. 2005;26(7):1015–1022. doi: 10.1016/j.neurobiolaging.2004.09.019. [DOI] [PubMed] [Google Scholar]
  • 81.Mondragon-Rodriguez S, Mena R, Binder LI, Smith MA, Perry G, Garcia-Sierra F. Conformational changes and cleavage of tau in Pick bodies parallel the early processing of tau found in Alzheimer pathology. Neuropathol Appl Neurobiol. 2008;34(1):62–75. doi: 10.1111/j.1365-2990.2007.00853.x. [DOI] [PubMed] [Google Scholar]
  • 82.Luna-Munoz J, Chavez-Macias L, Garcia-Sierra F, Mena R. Earliest stages of tau conformational changes are related to the appearance of a sequence of specific phospho-dependent tau epitopes in Alzheimer’s disease. J Alzheimers Dis. 2007;12(4):365–375. doi: 10.3233/jad-2007-12410. [DOI] [PubMed] [Google Scholar]
  • 83.Mondragon-Rodriguez S, Basurto-Islas G, Santa-Maria I, et al. Cleavage and conformational changes of tau protein follow phosphorylation during Alzheimer’s disease. Int J Exp Path. 2008;89(2):81–90. doi: 10.1111/j.1365-2613.2007.00568.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Garcia-Sierra F, Mondragon-Rodriguez S, Basurto-Islas G. Truncation of tau protein and its pathological significance in Alzheimer’s disease. J Alzheimers Dis. 2008;14(4):401–409. doi: 10.3233/jad-2008-14407. [DOI] [PubMed] [Google Scholar]
  • 85.Markesbery WR, Carney JM. Oxidative alterations in Alzheimer’s disease. Brain Pathol. 1999;9(1):133–146. doi: 10.1111/j.1750-3639.1999.tb00215.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Nunomura A, Perry G, Hirai K, et al. Neuronal RNA oxidation in Alzheimer’s disease and Down’s syndrome. Ann NY Acad Sci. 1999;89(3):362–364. doi: 10.1111/j.1749-6632.1999.tb07855.x. [DOI] [PubMed] [Google Scholar]
  • 87.Smith MA, Rudnicka-Nawrot M, Richey PL, et al. Carbonyl-related posttranslational modification of neurofilament protein in the neurofibrillary pathology of Alzheimer’s disease. J Neurochem. 1995;64(6):2660–2666. doi: 10.1046/j.1471-4159.1995.64062660.x. [DOI] [PubMed] [Google Scholar]
  • 88.Pappolla MA, Omar RA, Kim KS, Robakis NK. Immunohistochemical evidence of oxidative [corrected] stress in Alzheimer‘s disease. Am J Pathol. 1992;140(3):621–628. [PMC free article] [PubMed] [Google Scholar]
  • 89.Nunomura A, Chiba S, Lippa CF, et al. Neuronal RNA oxidation is a prominent feature of familial Alzheimer’s disease. Neurobiol Dis. 2004;17(1):108–113. doi: 10.1016/j.nbd.2004.06.003. [DOI] [PubMed] [Google Scholar]
  • 90.Nunomura A, Moreira PI, Takeda A, Smith MA, Perry G. Oxidative RNA damage and neurodegeneration. Curr Med Chem. 2007;14(28):2968–2975. doi: 10.2174/092986707782794078. [DOI] [PubMed] [Google Scholar]
  • 91.Nunomura A, Perry G, Pappolla MA, et al. RNA oxidation is a prominent feature of vulnerable neurons in Alzheimer’s disease. J Neurosci. 1999;19(6):1959–1964. doi: 10.1523/JNEUROSCI.19-06-01959.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Strocchi P, Pession A, Dozza B. Up-regulation of cDK5/p35 by oxidative stress in human neuroblastoma IMR-32 cells. J Cell Biochem. 2003;88(4):758–765. doi: 10.1002/jcb.10391. [DOI] [PubMed] [Google Scholar]
  • 93.Tamagno E, Robino G, Obbili A, et al. H2O2 and 4-hydroxynonenal mediate amyloid β-induced neuronal apoptosis by activating JNKs and p38MAPK. Exp Neurol. 2003;180(2):144–155. doi: 10.1016/s0014-4886(02)00059-6. [DOI] [PubMed] [Google Scholar]
  • 94.Webber KM, Smith MA, Lee HG, et al. Mitogen- and stress-activated protein kinase 1: convergence of the ERK and p38 pathways in Alzheimer’s disease. J Neurosci Res. 2005;79(4):554–560. doi: 10.1002/jnr.20380. [DOI] [PubMed] [Google Scholar]
  • 95.Munch G, Schinzel R, Loske C, et al. Alzheimer’s disease – synergistic effects of glucose deficit, oxidative stress and advanced glycation endproducts. J Neural Transm. 1998;105(4–5):439–461. doi: 10.1007/s007020050069. [DOI] [PubMed] [Google Scholar]
  • 96.Lu Z, Nie G, Li Y, et al. Overexpression of mitochondrial ferritin sensitizes cells to oxidative stress via an iron-mediated mechanism. Antioxid Redox Signal. 2009;11(8):1791–1803. doi: 10.1089/ars.2008.2306. [DOI] [PubMed] [Google Scholar]
  • 97.Bowser R, Smith MA. Cell cycle proteins in Alzheimer’s disease: plenty of wheels but no cycle. J Alzheimers Dis. 2002;4(3):249–254. doi: 10.3233/jad-2002-4316. [DOI] [PubMed] [Google Scholar]
  • 98.Mayer C, Grummt I. Cellular stress and nucleolar function. Cell Cycle. 2005;4(8):1036–1038. doi: 10.4161/cc.4.8.1925. [DOI] [PubMed] [Google Scholar]
  • 99.Tamagno E, Guglielmotto M, Aragno M, et al. Oxidative stress activates a positive feedback between the γ- and β-secretase cleavages of the β-amyloid precursor protein. J Neurochem. 2008;104(3):683–695. doi: 10.1111/j.1471-4159.2007.05072.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Yan SD, Yan SF, Chen X, et al. Non-enzymatically glycated tau in Alzheimer’s disease induces neuronal oxidant stress resulting in cytokine gene expression and release of amyloid β-peptide. Nat Med. 1995;1(7):693–699. doi: 10.1038/nm0795-693. [DOI] [PubMed] [Google Scholar]
  • 101.Nunomura A, Perry G, Pappolla MA, et al. Neuronal oxidative stress precedes amyloid-β deposition in Down syndrome. J Neuropathol Exp Neurol. 2000;59(11):1011–1017. doi: 10.1093/jnen/59.11.1011. [DOI] [PubMed] [Google Scholar]
  • 102.Nunomura A, Perry G, Aliev G, et al. Oxidative damage is the earliest event in Alzheimer disease. J Neuropathol Exp Neurol. 2001;60(8):759–767. doi: 10.1093/jnen/60.8.759. [DOI] [PubMed] [Google Scholar]
  • 103.Cuajungco MP, Goldstein LE, Nunomura A, et al. Evidence that the β-amyloid plaques of Alzheimer’s disease represent the redox-silencing and entombment of Aβ by zinc. J Biol Chem. 2000;275(26):19439–19442. doi: 10.1074/jbc.C000165200. [DOI] [PubMed] [Google Scholar]
  • 104.Zhu X, Rottkamp CA, Boux H, Takeda A, Perry G, Smith MA. Activation of p38 kinase links tau phosphorylation, oxidative stress, and cell cycle-related events in Alzheimer disease. J Neuropathol Exp Neurol. 2000;59(10):880–888. doi: 10.1093/jnen/59.10.880. [DOI] [PubMed] [Google Scholar]
  • 105.Zhu X, Castellani RJ, Takeda A, et al. Differential activation of neuronal ERK, JNK/SAPK and p38 in Alzheimer disease: the ‘two hit’ hypothesis. Mech Ageing Dev. 2001;123(1):39–46. doi: 10.1016/s0047-6374(01)00342-6. [DOI] [PubMed] [Google Scholar]
  • 106.Zhu X, Raina AK, Rottkamp CA, et al. Activation and redistribution of c-jun N-terminal kinase/stress activated protein kinase in degenerating neurons in Alzheimer’s disease. J Neurochem. 2001;76(2):435–441. doi: 10.1046/j.1471-4159.2001.00046.x. [DOI] [PubMed] [Google Scholar]
  • 107.Perry G, Nunomura A, Lucassen P, Lassmann H, Smith MA. Apoptosis and Alzheimer’s disease. Science. 1998;282(5392):1268–1269. doi: 10.1126/science.282.5392.1265h. [DOI] [PubMed] [Google Scholar]
  • 108.Mecocci P, MacGarvey U, Beal MF. Oxidative damage to mitochondrial DNA is increased in Alzheimer’s disease. Ann Neurol. 1994;36(5):747–751. doi: 10.1002/ana.410360510. [DOI] [PubMed] [Google Scholar]
  • 109.Holley AK, St Clair DK. Watching the watcher: regulation of p53 by mitochondria. Future Oncol. 2009;5(1):117–130. doi: 10.2217/14796694.5.1.117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Zhu X, Raina AK, Boux H, Simmons ZL, Takeda A, Smith MA. Activation of oncogenic pathways in degenerating neurons in Alzheimer disease. Int J Dev Neurosci. 2000;18(4–5):433–437. doi: 10.1016/s0736-5748(00)00010-1. [DOI] [PubMed] [Google Scholar]
  • 111.Zhu X, Raina AK, Smith MA. Cell cycle events in neurons. Proliferation or death? Am J Pathol. 1999;155(2):327–329. doi: 10.1016/S0002-9440(10)65127-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.McShea A, Lee HG, Petersen RB, et al. Neuronal cell cycle re-entry mediates Alzheimer disease-type changes. Biochim Biophys Acta. 2007;1772(4):467–472. doi: 10.1016/j.bbadis.2006.09.010. [DOI] [PubMed] [Google Scholar]
  • 113.Lee HG, Casadesus G, Nunomura A, et al. The neuronal expression of MYC causes a neurodegenerative phenotype in a novel transgenic mouse. Am J Pathol. 2009;174(3):891–897. doi: 10.2353/ajpath.2009.080583. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Cuadra G, Giacobini E. Effects of cholinesterase inhibitors and clonidine coadministration on rat cortex neurotransmitters in vivo. J Pharmacol Exp Ther. 1995;275(1):228–236. [PubMed] [Google Scholar]
  • 115.Pepeu G, Giovannini MG. Cholinesterase inhibitors and memory. Chem Biol Interact. 2009 doi: 10.1016/j.cbi.2009.11.018. (Epub ahead of print) [DOI] [PubMed] [Google Scholar]
  • 116.Gandy S, Greengard P. Amyloidogenesis in Alzheimer’s disease: some possible therapeutic opportunities. Trends Pharmacol Sci. 1992;13(3):108–113. doi: 10.1016/0165-6147(92)90039-9. [DOI] [PubMed] [Google Scholar]
  • 117.Yamada K, Yabuki C, Seubert P, et al. Aβimmunotherapy: intracerebral sequestration of Aβ by an anti-Aβ monoclonal antibody 266 with high affinity to soluble Aβ. J Neurosci. 2009;29(36):11393–11398. doi: 10.1523/JNEUROSCI.2021-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Lambracht-Washington D, Qu BX, Fu M, Eagar TN, Stuve O, Rosenberg RN. DNA β-amyloid1–42 trimer immunization for Alzheimer disease in a wild-type mouse model. JAMA. 2009;302(16):1796–1802. doi: 10.1001/jama.2009.1547. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Town T. Alternative Ab immunotherapy approaches for Alzheimer’s disease. CNS Neurol Disord Drug Targets. 2009;8(2):114–127. doi: 10.2174/187152709787847306. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Santa-Maria I, Hernandez F, Del Rio J, Moreno FJ, Avila J. Tramiprosate, a drug of potential interest for the treatment of Alzheimer’s disease, promotes an abnormal aggregation of tau. Mol Neurodegener. 2007;2:17. doi: 10.1186/1750-1326-2-17. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Middleton LE, Yaffe K. Promising strategies for the prevention of dementia. Arch Neurol. 2009;66(10):1210–1215. doi: 10.1001/archneurol.2009.201. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Marlatt MW, Lucassen PJ, Perry G, Smith MA, Zhu X. Alzheimer’s disease: cerebrovascular dysfunction, oxidative stress, and advanced clinical therapies. J Alzheimers Dis. 2008;15(2):199–210. doi: 10.3233/jad-2008-15206. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES