Next Article in Journal
Nanotechnology Facilitated Cultured Neuronal Network and Its Applications
Next Article in Special Issue
Discovery of a Metabolic Signature Predisposing High Risk Patients with Mild Cognitive Impairment to Converting to Alzheimer’s Disease
Previous Article in Journal
Cold Atmospheric Pressure Plasma-Activated Medium Induces Selective Cell Death in Human Hepatocellular Carcinoma Cells Independently of Singlet Oxygen, Hydrogen Peroxide, Nitric Oxide and Nitrite/Nitrate
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Alzheimer’s Disease Animal Models: Elucidation of Biomarkers and Therapeutic Approaches for Cognitive Impairment

1
Department of Neuropsychopharmacology and Hospital Pharmacy, Nagoya University Graduate School of Medicine, Nagoya 466-8560, Japan
2
Medical Interactive Research and Academia Industry Collaboration Center, Research Institute of Environmental Medicine, Nagoya University, Nagoya 464-8601, Japan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(11), 5549; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22115549
Submission received: 29 April 2021 / Accepted: 21 May 2021 / Published: 24 May 2021
(This article belongs to the Special Issue Molecular Mechanisms of Dementia—Application to Its Biomarkers)

Abstract

:
Alzheimer’s disease (AD) is an age-related and progressive neurodegenerative disorder. It is widely accepted that AD is mainly caused by the accumulation of extracellular amyloid β (Aβ) and intracellular neurofibrillary tau tangles. Aβ begins to accumulate years before the onset of cognitive impairment, suggesting that the benefit of currently available interventions would be greater if they were initiated in the early phases of AD. To understand the mechanisms of AD pathogenesis, various transgenic mouse models with an accelerated accumulation of Aβ and tau tangles have been developed. However, none of these models exhibit all pathologies present in human AD. To overcome these undesirable phenotypes, APP knock-in mice, which were presented with touchscreen-based tasks, were developed to better evaluate the efficacy of candidate therapeutics in mouse models of early-stage AD. This review assesses several AD mouse models from the aspect of biomarkers and cognitive impairment and discusses their potential as tools to provide novel AD therapeutic approaches.

1. Introduction

Dementia is defined as a syndrome, caused by various brain disorders, that affects memory, thinking, behavior, and the ability to perform daily activities [1]. Worldwide, there are approximately 50 million people living with dementia. Alzheimer’s disease (AD) accounts for 60–80% of these cases [2]. AD is an irreversible and progressive neurodegenerative disorder that presents with cognitive impairment and memory loss as the primary clinical symptoms. The disease is mainly caused by the accumulation of extracellular amyloid β (Aβ) and intracellular neurofibrillary tau tangles [3]. Structural brain changes are thought to begin 20 years or more before the onset of clinical symptoms [4]. The disease has several phases, which include a long preclinical phase with no clinical symptoms, a mild cognitive impairment phase, and a disease phase [5]. The length of each phase is affected by the patient’s age, genetics, gender, and other factors [6]. Although AD has been studied for over 100 years, fundamental treatment strategies for the disease remain underdeveloped. It is therefore important to develop new therapeutic medications for AD patients that are more effective than the existing treatments.
Rodent models are evaluated using behavioral tests such as the Morris water maze (MWM), the radial maze, the Y-maze, the T-maze, fear conditioning (FC), and novel object recognition (NOR) tests. These tests play a crucial role as indicators of learning, memory, and cognitive functions, which correspond to the late phases of cognitive deficits in AD patients [7,8]. This review summarizes the characteristics of several representative AD mouse models that were generated based upon the amyloid cascade hypothesis. In addition, we assess these models in regard to changes in biomarkers, pathologies, and behavior, and we discuss their potential as useful tools to provide therapeutic approaches in AD. In conjunction with this analysis, we introduce several approaches to develop new therapeutic strategies for AD.

2. Amyloid Cascade Hypothesis

The amyloid cascade hypothesis was first proposed in 1991 by Hardy and Allsop [9]. They identified a pathogenic mutation in the Aβ precursor protein (APP) gene on chromosome 21 in familial AD (FAD) patients. The mutation causes APP mismetabolism and Aβ deposition, which are primary events in AD disease progression. They further suggested that the pathology of AD is caused by abnormal fibrous protein deposits, including senile plaques and neurofibrillary tangles (NFTs), and amyloid deposits on the walls of cerebral blood vessels. Senile plaques and cerebrovascular amyloids are caused by the deposition of extracellular Aβ. Aβ first isolated from the meningeal vessels of AD patients by Glenner and Wong in 1984 [10], is a 39–43 residue protein derived from multiple proteolytic cleavages of APP. APP was first cloned and sequenced in 1987 [11,12], is 695 amino acid residues in length, and may function as a glycosylated receptor on the cell surface [13]. APP is degraded by the proteolytic enzymes α-, β-, and γ-secretase in two processing pathways [14]: Aβ–non-producing and Aβ-producing. In the Aβ–non-producing pathway, APP is hydrolyzed by α-secretase and then by γ-secretase, which has presenilin 1 (PSEN1) as the core active subunit. ADAM9, ADAM10, and ADAM17 may also have α-secretase activity [15]. Importantly, these cleavage events do not lead to Aβ deposition in the brain because this pathway does not produce insoluble Aβ. By contrast, in the Aβ-producing pathway, APP is hydrolyzed by β-secretase (BACE1) followed by γ-secretase, which leads to the production of insoluble Aβ [14]. The levels of insoluble Aβ may play an important role in the onset and progression of AD [16]. As APP hydrolysis mainly occurs through the Aβ–non-producing pathway, insoluble Aβ protein is not produced under normal conditions. The small amount of APP hydrolyzed via the Aβ-producing pathway is usually eliminated by the immune system. Aβ may function as a neurotrophic factor for differentiating neurons, but the high concentrations of Aβ in AD patients cause neuronal death [17]. Moreover, cholinergic and noradrenergic neurons are highly sensitive to amyloid toxicity, resulting in a decrease in these neurons in the preclinical phase, well before the appearance of amyloid plaques and cognitive deficits [18,19]. In FAD patients harboring APP mutations near the BACE1 cleavage site, APP is likely hydrolyzed by the Aβ-producing pathway [20,21,22], resulting in excessive accumulation of insoluble Aβ and the eventual development of AD. Additionally, two Aβ peptides are formed by BACE1 and γ-secretase cleavage events: a peptide terminating at amino acid residue 40 (Aβ40) and a longer form ending at residue 42 (Aβ42) [16,23]. Aβ42 is more likely to misfold and aggregate than Aβ40, suggesting that Aβ42 is more neurotoxic [24]. Aβ42 plasma levels are elevated in AD patients, suggesting that this increase is associated with the development of AD [25].

3. Tau Propagation Hypothesis

The formation of intracellular NFTs, derived from microtubule-associated tau, is also a hallmark of AD pathology. The tau propagation hypothesis was introduced in 2009 [26]. Tau, a 352 residue thermostable protein, was identified in 1975 [27] and cloned and sequenced in 1988 [28]. Tau stabilizes microtubules, thereby promoting the polymerization of tubulin into microtubules [29]. Although phosphorylated tau maintains the cytoskeleton, neuronal tau may regulate the stability of axonal microtubules and eventually transports signal transduction–associated proteins through the microtubules. In the human brain, six predominant tau isoforms are formed by selective splicing [30]. These isoforms are distinguished by the presence or absence of two N-terminal inserts (exon 2 and 3 inclusion) and the presence of either three or four imperfect arginine repeats (3R or 4R) in the microtubule binding domain at the C-terminus (exon 10 inclusion) [31]. These observations suggest that a balanced tau isoform ratio is necessary for maintaining normal brain function in humans, and an imbalance in these ratios would therefore lead to tau aggregation into NFTs. In humans, tau pathology is observed initially in a specific area of the brain and then spreads to other areas [32]. Although tau has 85 potential serine, threonine, and tyrosine phosphorylation sites [33], abnormal phosphorylation at approximately 45 different sites is related to AD pathology [34]. Excessive or abnormal phosphorylation may be caused by alterations in the activities of various kinases or phosphatases that target tau. Abnormally phosphorylated tau forms anti-helical fibrils and depolymerizes microtubes, which is followed by the formation of insoluble NFTs [35]. The insoluble NFTs may impair cytoplasmic function and interfere with axonal transport between neurons, thereby inducing cell death. The molecular and cellular mechanisms responsible for the formation of tau lesions remain unclear.

4. Biomarkers of AD

Based on the amyloid cascade and tau propagation hypotheses, the measurement of Aβ40, Aβ42, and tau in cerebrospinal fluid (CSF) was established as diagnostic for AD biomarkers. A decrease in CSF Aβ42 levels [36], occurring before AD onset [37], and elevated tau levels [38,39,40] were observed in AD patients. However, an increase in CSF tau levels was also detected in other neurological diseases, including psychiatric disorders [38,39,40] and normal aging [38,39]. For this reason, it was deemed advantageous to combine the measurement of Aβ40, Aβ42, and tau in CSF to confirm AD in patients, and it was hoped that these measurements would be a useful biological marker for AD therapeutic strategies [41]. In addition, Aβ oligomers in CSF may play a critical role in the pathogenesis and progression of AD [42,43]. However, the measurement of CSF biomarkers is a highly invasive test, and the physical burden on the elderly is particularly high. Positron emission technology (PET) using 11C-labeled Pittsburgh Compound B (11C-PIB) to visualize amyloid accumulation in the brain of AD patients was developed [44,45]. Aβ PET radiotracers enable detection of early-stage AD pathology [46]. Consequently, amyloid PET is widely used to diagnose preclinical AD. 18F-Florbetapir, approved by the United States Food and Drug Administration (FDA), was the first 18F-labeled tracer developed to detect Aβ in 18F-fluorodeoxyglucose (18F-FDG) PET [47]. Although 18F-Florbetapir is the most widely used Aβ tracer, 11C-PIB has a higher binding affinity for amyloid and a shorter half-life. As an amyloid PET scan can only be performed in facilities that are equipped with a cyclotron and a device for preparing labeled compounds, this diagnostic is often cost-prohibitive. Recently, PET imaging that detects tau accumulation [48,49] or microglia activation [50,51] was also developed. Amyloid PET scans are expensive and CSF collection is invasive, therefore the development of tests for measuring AD biomarkers from blood samples was highly anticipated. Although it is difficult to detect biomarkers in human blood [52,53,54], mass spectrometry [55,56,57] and enzyme-linked immunosorbent assay [58] successfully detect amyloid in peripheral blood. Determination of the alteration of Aβ secondary structure in blood plasma by an immune infrared sensor is also possible, suggesting that this change could be used as a blood biomarker for severe AD stages [59]. Additionally, the usefulness of measuring tau [60,61,62], Interleukin-8 [63], and neurofilament light [64] levels in blood plasma are confirmed in preclinical AD. The analysis of microRNAs in blood was also shown to be effective for prospective AD risk prediction, suggesting that microRNAs may be of practical clinical use as biomarkers of AD [65]. Moreover, signal transduction proteins, such as insulin-like growth factor (IGF) 1 and IGF-binding protein 3, may predict cognitive decline [66]. Aβ clearance proteins, including apolipoprotein (APO) A-I, complement protein C3, and transthyretin in the serum, may also be potential biomarkers for mild cognitive impairment evaluation [67]. However, it is difficult to unify the analysis techniques for these biomarkers worldwide, and these biomarker tests are not yet widespread. If AD biomarkers from blood samples can be established and used for pre-screening before expensive PET imaging, it would greatly decrease the cost of developing therapeutic or preventive drugs, consequently accelerating their development. An overview of these biomarkers can be found in Table 1.

5. Biomarkers and Amyloid Cascade Hypothesis-Related Animal Models

In FAD patients, accumulating evidence indicates that Aβ begins to deposit extracellularly 20–30 years before the onset of cognitive and memory impairments and before secondary tau appears in neurons. For this reason, it is widely believed that the amyloid cascade hypothesis is the main working hypothesis in AD. To investigate AD pathology, molecular mechanisms, and cognitive function, APP and PSEN gene mutations were identified in FAD patients [20,68]. Transgenic (Tg) mice that overexpresses the APP gene mutation associated with FAD (APP-Tg mice) were generated as AD mouse models [2]. Over the past 20 years, many additional Tg mice, carrying multiple mutations including the APP and PSEN familial gene mutations, were developed. Molecular biological studies in these models indicate that several genes and their mutations play a role in the development of early-onset AD. In addition to the APP, PSEN1, and PSEN2 genes, more than 20 genetic risk loci for AD were identified [69]. Among these genes, a mutation in apolipoprotein E (APOE) and a rare variant of the triggering receptor expressed on myeloid cells 2 (TREM2) are the most common genetic risk factors for late-onset AD. Below, we will briefly summarize the advantages and limitations of several popular Tg mice as AD mouse models. An overview of the timeline for amyloid, phosphorylated tau, and NFT pathologies observed in these APP-Tg mice can be found in Table 2.

5.1. Animal Models for Early-Onset AD

5.1.1. Tg2576

The APPSWE (Tg2576) mice overexpress the 695–amino acid isoform of human APP protein [containing the Swedish mutation (K670N and M671L, and KM670/671NL)] under the control of the hamster prion protein (PrP) promoter [70]. Tg2576 mice show a 5-fold increase in Aβ40 and a 14-fold increase in Aβ42/43 concentrations in the old mice compared to the young mice. Inflammatory markers, such as elevated levels of cytokines and microglia activation, are observed in aged mice [71]. In behavioral tests, these mice also exhibited learning and memory impairments in spatial reference and alternation tasks at 9–10 months of age, although no significant differences were found at 3 months of age [70]. Additionally, Tg2576 mice exhibit learning- and memory-associated abnormalities in the T-maze alternation task [72] and hippocampus-dependent fear memory [72,73] and amygdala-dependent cued fear learning ability in the FC test [74], suggesting that these mice have a widespread deficit in cognitive functions. The behavioral and pathological features of Tg2576 mice resemble those observed in human AD. Alterations in adult brain neurogenesis were reported for Tg2576 mice and human AD patients [75]. In contrast to the aforementioned similarities between Tg2576 mice and human AD patients, there are several key differences. Although widespread neuronal cell loss and NFT pathology are observed in the cortex and hippocampus in advanced-stage AD patients [76], this is not recapitulated in the Tg2576 mice [77]. Brain dysfunction occurs prior to amyloid deposition in this mouse model, suggesting that the symptoms observed could be due to soluble Aβ species.

5.1.2. APP23

The APP23 mice overexpress the human APP protein with the Swedish double mutations, which combine KM670/671NL with the V717I mutation, under the control of the mouse Thy1 promoter [78]. As this mouse model has the same Swedish mutations found in the Tg2576 mouse, the neuropathological and behavioral phenotypes of APP23 mice resemble those of Tg2576. Amyloid plaques in this model are observed at 6 months of age [78], followed by neuritic and synaptic degenerations and abnormal tau phosphorylation induced by Aβ plaque deposition with aging. Neuronal loss is found in the CA1 region of the hippocampus [79] and neocortex [80] in the brains of these mice. Similar to AD patients and Tg2576 mice, APP23 mice exhibit amyloid angiopathy in the cerebral vessels [81]. Additionally, APP23 Tg mice display age-related abnormal phenotypes associated with cognitive function in various behavioral tests [82,83,84,85,86]. These phenotypes in APP23 mice are similar to all symptoms observed in AD patients.

5.1.3. PDAPP

The PDAPP mice overexpress the human APP protein with the Indiana mutation (V717F) under the control of a platelet-derived growth factor-β promoter [87,88]. Similar to AD patients, PDAPP mice show an accumulation of amyloid plaques beginning at 6–9 months of age, resulting in glial activation, dystrophic neurites, gliosis, and loss of synaptic and dendritic density in the hippocampus. PDAPP mice express high levels of human APP, more than 10-fold higher than endogenous murine APP. Unlike AD patients, PDAPP mice do not have NFTs or neuronal loss, although immunoreactivity of phosphorylated tau is observed in dystrophic neurites [89]. The PDAPP and Tg2576 mice differ in neuropathology. Tg2576 mice show a progressive increase in both Aβ40 and Aβ42 levels [70], whereas PDAPP mice do not have elevated levels of Aβ42. A number of amyloid deposits in dense-cored plaques were observed in Tg2576 mice, whereas PDAPP mice have only a few diffuse deposits [87,88]. Moreover, unlike Tg2576 mice, giant plaques and vascular amyloid depositions are also largely absent in PDAPP mice. This mouse model has cognitive and memory deficits at a young age in a variety of behavioral tests prior to Aβ plaque deposition [90,91,92].

5.1.4. TgCRND8

The TgCRND8 mice overexpress a double mutant form of the human APP protein (KM670/671NL and V717F) under the control of the PrP promoter [93]. This mouse model has thioflavin S-positive Aβ deposits at 3 months of age that closely resemble those seen in AD. Dense-cored plaques and selective neuronal loss are detected at 5 months of age [93], followed by the accumulation of hyperphosphorylated tau at 7–12 months of age in the neocortex and in the dentate gyrus (DG), CA1, and CA3 regions of the hippocampus [94]. No deposition of NFTs is observed in this model [93]. Neuromorphological abnormalities do not appear in this mouse model until 6 months of age [95]. Age-related behavioral impairments in TgCRND8 mice are observed starting at 3 months of age in spatial working and reference memory tasks including the Barnes maze, MWM, NOR, and FC tests [93,96,97,98,99,100]. Therefore, this mouse model, as well as Tg2576, could be potentially powerful models for developing therapeutic drugs to treat early-onset AD. However, Aβ deposits and dystrophic neurites in TgCRND8 and Tg2576 mice are similar to those noted in human pathological aging but not to those in AD [101].

5.1.5. APPPS1

The APPPS1 mice overexpress human APP (with the KM670/671NL mutation) and PSEN1 (with the L166P mutation), both under the control of the mouse Thy1 promoter [102]. This model has an impairment in amyloid protein processing, resulting in elevated Aβ42 levels at 2–3 months of age [102]. Cerebral amyloidosis is observed starting at 6–8 weeks-old. Formation of NFTs is not observed in this model, although hyperphosphorylated tau-positive neurites are detected at 8 months of age. Various pathological abnormalities, including glial activation [103], dystrophic neurites, gliosis, and loss of synaptic and dendritic density [104], are induced by an age-related accumulation of amyloid plaques in these mice. Although neuronal loss in the neocortex is not observed at 8 months of age, loss was seen in the DG of the hippocampus and other subregions in older mice [103]. These mice show spatial learning and memory impairments at 6–8 months of age in a food-rewarded four-arm spatial maze [102], MWM, and radial maze tests [105,106]. Furthermore, at 120–250 days, this model exhibits memory impairments in NOR [107] and Barns maze tests [108]. Interestingly, impairments in reversal learning in 6–9-month-old APPPS1 mice were confirmed by touchscreen visual discrimination tasks [109]. Due to the early onset of amyloidosis and the stable genetic background of this line, this mouse model is well-suited for studying the pathology of AD amyloidosis and investigating novel therapeutic strategies.

5.1.6. 5XFAD

5XFAD mice overexpress the human APP and PSEN1 proteins with a total of five AD-linked mutations under the control of the mouse Thy1.2 promoter [110]. The mutations are the Swedish (K670N/M671L), Florida (I716V), and London (V717I) mutations in APP, and the M146L and L286V mutations in PSEN1. Intracellular Aβ starts to accumulate in 1.5-month-old mice, and extracellular Aβ deposition appears in 2-month-old mice [110]. 5XFAD mice accumulate more Aβ42 in the cerebrum than Aβ40, suggesting that the five FAD mutations cumulatively affect Aβ42 production. The levels of Aβ42 in this model are far higher than those found in Tg2576. However, tau hyperphosphorylation and formation of NFTs are not observed in 5XFAD mice. Astrogliosis and microgliosis present at 2 months of age, indicating that neuroinflammation occurs early in this model. Moreover, 5XFAD mice show a progressive loss of neurons, which correlates strongly with both intraneuronal Aβ42 levels and caspase-3 activation [111]. These mice exhibit progressive cognitive and memory impairments in Y-maze [112], NOR [113], FC [114], and MWM tests [112,113,114,115,116]. This mouse model could be a useful model, and it reproduces the pathology of human AD, similar to the Tg2576, APP23, and APPPS1 models.

5.1.7. 3×Tg-AD

The Triple Tg (3×Tg-AD) mice overexpress the human APP protein with the Swedish mutation, the mutant PSEN1 protein, and the mutant human microtubule-associated protein tau (APPK670N,M671L, PSEN1M146V, and MAPTP301L) [117]. APPK670N,M671L and PSEN1M146V are controlled by the mouse Thy1.2 promoter. This model progressively accumulates Aβ plaques, hyperphosphorylated tau, and NFTs. Before these deposits form, synaptic dysfunction and long-term potentiation (LTP) deficits occur in this model. Intracellular Aβ deposits appear in the brain at 3 months of age in the frontal cortex and at 6 months of age in the CA1 region of the hippocampus [117] and the amygdala [118], followed by extracellular Aβ deposition progresses with aging [117]. These mice exhibit elevated levels of Aβ40 and Aβ42 with aging. Formation of intracellular NFTs, composed of hyperphosphorylated tau protein, are observed in the hippocampus at 12 months of age but not at 6 months of age. In addition, 3×Tg mice develop age-dependent synaptic dysfunction [117] and loss of tyrosine hydroxylase–positive neurons at the coeruleus locus [119]. This model does not present with the neuronal loss in the hippocampus observed in AD patients. Microglia activation is observed at 7 months of age [120]. Unlike human AD patients, no difference is observed between wild-type and 3×Tg-AD mice in white matter densities measured using magnetic resonance imaging (MRI) [121]. Behaviorally, young 3×Tg-AD mice exhibit an impairment in learning and memory deficits in MWM; these changes correlate with the accumulation of intraneuronal Aβ in the hippocampus and amygdala [118]. Additionally, this mouse model exhibits progressive cognitive and memory impairments in various behavioral tests. [118,122,123,124,125,126,127]. As the tau P301L mutation in 3×Tg-AD mice is the causative gene mutation in frontotemporal dementia but not in AD, tau pathology would differ from that observed in human AD. This model has natural tau mutations in the tau P301L background, and therefore would not express proteins that follow the amyloid cascade hypothesis. Consequently, 3×Tg-AD mice do not seem to be a mouse model that faithfully reflects AD pathology.

5.2. Animal Models for Late-Onset AD

5.2.1. APOE

APOE is a lipid metabolism–associated gene that is localized to senile plaques, vascular amyloid deposits, and NFTs in AD. The APOE gene is located on chromosome 19q13.2 and has three alleles, ε2, ε3, and ε4, which are present at frequencies of 8.4%, 77.9%, and 13.7%, respectively [128]. The differences between APOE2 (Cys112, Cys158), APOE3 (Cys112, Arg158), and APOE4 (Arg112, Arg158) are limited to amino acid residues 112 and 158 [129]. The presence of the APOE ε4 allele is the strongest genetic risk factor for AD [129]. APOE4 is also associated with late-onset AD and is an important susceptibility marker for AD [129]. Additionally, APOE4 increases the neurotoxicity of Aβ, tau hyperphosphorylation, and NFTs and influences the timing and amount of amyloid deposition in the human brain [129]. The neuronal glycoprotein reelin may protect synapses against toxic Aβ through APOE receptors [130]. Therefore, APOE is a potential target for AD therapeutics. To investigate these functions, mice harboring modifications in the APOE gene [APOE-knock-in (APOE-KI), APOE knockout (APOE KO), and APOE-targeted replacement (APOE-TR) mice] were generated [129]. In particular, human APOE4 KO and APOE4-TR mice exhibit neuronal deficits and cognitive impairments [129,131,132,133,134]. To investigate the effects of APOE on APP-induced neuropathology, APOE-KI/KO/TR mice were bred with APP-Tg mice [131,135,136,137]. APOE KO/PDAPP mice exhibited a dramatic reduction in amyloid plaque deposition at 6 months of age in the cerebral cortex and hippocampus [135]. Older human APOE4-TR/PDAPP mice increased the levels of insoluble APOE protein and Aβ in the brain at the same time, which decreased that of soluble APOE protein [131]. On the other hand, APOE4-KI/5XFAD mice exhibit delayed amyloid plaque deposition, although 5XFAD mice develop amyloid plaques by 2 months of age [137]. The delay in amyloid plaque accumulation is also observed in APOE4-KI/Tg2576 mice [136]. These observations suggest that APOE modulates Aβ deposition. Mice harboring APOE modifications crossed with APP-Tg mice could be useful tools for studying AD pathology, but these models are still under development.

5.2.2. TREM2

TREM2, like APOE, is also a strong genetic risk factor for AD. TREM2 is located on human chromosome 6p21.1 and in the IgV domain on mouse chromosome 17 [138]. In regions such as the hippocampus, spinal cord, and white matter in aged mice and elderly humans, TREM2 is highly expressed in microglia and myeloid cells [139,140]. A number of TREM2 variants were identified as risk factors for late-onset AD [141]. Most of the variants affect the phagocytosis, maturation, and ligand affinity of TREM [142]. In the AD brain, microglia are thought to play a neuroprotective role by promoting Aβ phagocytosis, degradation, and clearance and by decreasing tau propagation. To investigate these functions, various TREM2 KO mice were generated. TREM2 KO mice exhibit reduced microglial activation, microgliosis, and phagocyte levels [143], suggesting that TREM2 modulates the inflammatory response and phagocytosis in microglia. Expression of TREM2 increases in AD patients [144,145]. Adeno-associated virus-mediated soluble TREM2 (sTREM2) expression reduces amyloid plaque load and rescues spatial memory and LTP deficits in 5XFAD mice [116]. sTREM2 enhances microglial proliferation, migration, clustering in the vicinity of amyloid plaques, and the uptake and degradation of Aβ, suggesting that sTREM2 may protect against AD pathology [116]. To further understand the role of microglia in AD, TREM2 KO mice were crossed with APP-Tg mice [146,147,148,149]. TREM2 KO/APPPS1 mice exhibit a reduction in inflammation and the accumulation of amyloid and tau [146], suggesting that TREM2 is involved in AD pathology. Hence, mice harboring TREM2 modifications crossed with APP-Tg mice could be used to investigate novel AD therapeutic approaches that target the microglia.
Table 2. Summary of representative APP-Tg and APP-KI mice in AD model mice.
Table 2. Summary of representative APP-Tg and APP-KI mice in AD model mice.
Mouse LinePromoterTransgene MutationAmyloid Plague-DepositsHyperphosphorylated TauNFTsReference
Tg2576Hamster Prion ProteinAPP Swedish mutation11–13 monthsNot detectedNot detected[70]
APP23Mouse Thy1APP Swedish mutation6 months6 monthsNot detected[78]
PDAPPPlatelet-derived growth factor-βAPP Indiana mutation6–9 months14 monthsNot detected[88,89]
TgCRND8Hamster Prion ProteinAPP Swedish + Indiana mutations3–5 months7–12 monthsNot detected[93,94]
APPPS1Mouse Thy1 (APP, PS1)APP Swedish + PS1 L166P mutations2–3 months8 monthsNot detected[102]
5XFADMouse Thy1.2 (APP, PS1)APP Swedish + Florida + London + PS1 M146V + L286V mutations1.5 monthsNot detectedNot detected[110]
3×Tg-ADMouse Thy1.2 (APP, Tau) and endogenous (PS1)APP Swedish + PS1 M146V + Tau P301L mutations3–6 months12 months12 months[117]
APP-KIEndogenous APPAPP Swedish + Iberian + Arctic mutations2 monthsNot detectedNot detected[150]

6. The Next Generation of Mouse Models and New Approaches for AD Treatment

Although APP-Tg mice have been used over the past 5 years to develop new AD therapeutic strategies using behavioral tests related to cognition and memory formation (Table 3), these mouse models have intrinsic problems that induce artificial phenotypes. In these models, some APP fragments, including Aβ, are over-produced. For example, APP23 mice have high levels of APP intracellular domain (AICD) [151] and C-terminal fragment β (CTF-β) [152]. These APP fragments may be involved in various functions including the control of Aβ degradation, cell death, and γ-secretase activity. Unlike AD patients, Tg2576 and APP23 mice exhibit the accumulation of different Aβ species, including Aβ40 and Aβ42. Aβ42 is the predominant species accumulated in the brain of AD patients. There are also differences in the generation of hyperphosphorylated tau among APP-Tg mice, although APP-Tg mice, with the exception of 3×Tg-AD mice, just overexpress the mutant human APP protein. These differences could be caused by overexpression of APP transgenes with artificial mutations [150].
To overcome these undesired problems, an APP gene knock-in (APP-KI) mouse with the Swedish (KM670/671NL), Beyreuther/Iberian (I716F), and Arctic mutations was generated [150]. This model overproduces Aβ42 without overexpressing APP. APP-KI mice exhibit excessive Aβ deposition in the cortex and hippocampus with age. Additionally, microgliosis and astrocytosis are observed at 9 months of age. Moreover, synaptic alternation and memory impairment, similar to what is seen in AD patients, are observed in these mice. On the other hand, as with other AD models, this model does not have tau pathology, NFTs, neurodegeneration, or massive neuron loss [153]. Therefore, this mouse model should be used to study preclinical AD. Various classical behavioral tasks to evaluate cognitive and memory functions at the early stage were performed with this model [154,155,156,157]. Behavioral impairments were not observed in young APP-KI mice, suggesting that these tests are unsuitable for evaluating the preclinical and early-onset stages of AD.
In recent years, it has become clear that early detection of AD is important for effective interventions [158]. Clinical studies demonstrate that early treatment initiation with the currently existing medications provides a more effective clinical benefit than later initiation of such therapy [159,160]. To better evaluate cognitive function in psychiatric and neurodegenerative disorders, new technology using touchscreen-based tasks was developed [161,162,163]. This technology assesses attention, learning, and memory in animals and humans. We showed that touchscreen-based tasks can detect cognitive impairment in APP-KI mice at 4–5 months of age [164]. This was the first report to show the benefit of these methods in detecting the early stage of cognitive impairment in an AD-linked APP-KI mouse model. Therefore, this advanced technology would detect preclinical AD-like behaviors in AD mouse models and would aid in the search for new therapeutic approaches to prevent AD progression.
Although various studies have focused on the relationship between AD and microglia, Sobue et al. [165] recently performed RNA sequencing using magnetic-activated cell sorting to analyze the microglia of three neurodegenerative disease-associated model mice: APP-KI with amyloid pathology, rTg4510 with tauopathy, and SOD1G93A with motor neuron disease. RNA sequencing was also used to analyze preclinical AD human precuneus, and the results were compared to those of a control group. Interestingly, the loss of unique microglial homeostatic genes in the progression of AD correlates with the severity of neurodegeneration. Moreover, RNA sequencing results for the human precuneus samples led us to conclude that amyloid pathology in early-onset AD induces loss of microglia and oligodendrocyte function. Further advances in these beneficial tools are expected in the future.
Table 3. Summary of reference with learning- and memory-associated behavioral tests in popular AD model mice over the past 5 years.
Table 3. Summary of reference with learning- and memory-associated behavioral tests in popular AD model mice over the past 5 years.
Behavioral TestsTest SignificanceSignificant Difference
Tg2576 APP23PDAPPTgCRND8 APPPS15XFAD3xTg-ADAPP-KI
Y-mazeShort-term working memory Yes: [166]
No: [125]
Yes: [112]Yes: [125,126]No: [156]
Hole board Reference and working memory Yes: [167]
No: [167]
Open-field foraging taskWorking memory Yes: [168]
Object in place taskSpatial recognition memory No: [169]
Object place recognitionShort-term memoryYes: [125]Yes: [125]
Object-place association taskRecognition memoryYes: [98]
Spatial object location task Recognition memoryYes: [98]
Spatial memory No: [170]
Novel object recognitionRecognition memoryYes: [171]
No: [169]
Yes: [86]Yes: [98,100]Yes: [108]Yes: [113]Yes: [122,123]No: [156]
Hippocampal-dependent episodic memoryYes: [125] Yes: [125]
Social preference social novelty Social memoryNo: [157]
Fear conditioning Associative memory No: [172]No: [173]
Contextual fear memory Yes: [99]Yes: [170]
Tone-cued fear memory No: [170]
Non-hippocampal-dependent auditory fear memory Yes: [99]
Fear learning No: [155]
Fear conditioning context discriminationContext discrimination learning and memory Yes: [73]
Passive avoidanceContextual learning and memoryYes: [122]
Learning and memoryYes: [126]
Eight-arm radial maze Working memoryYes: [84]
Spatial learning and memory Yes: [100]
Banes maze Spatial learning and memory Yes: [108]Yes: [155]
Spatial navigation memoryYes: [98]
Morris water mazeSpatial learning and memoryYes: [124,166,171,174] No: [175]Yes: [85] No: [86]Yes: [99]Yes: [112,113,115]Yes: [122,124,125,126,127]
Spatial recognition memoryNo: [174]
Spatial reference memory No: [156,157,164]
Spatial reversal learningFlexibility and impulse controlNo: [155]
Location discriminationPattern separationYes: [164]
Different object –location paired-associate learningPaired-associative memoryYes: [164]
Visual discrimination, reversal learningCognitive flexibilityYes: [109]No: [164]

7. Pharmacological Interventions in AD Animal Models

Over the years, various amyloid cascade hypothesis–related AD model mice were generated to develop new therapeutic drugs for AD treatment. Although the acetylcholinesterase inhibitors donepezil, galantamine, and rivastigmine, and the N-methyl-D-aspartate receptor inhibitor memantine are FDA-approved therapeutics, these drugs are only able to delay the progression of AD. All of the approved drugs ameliorate cognitive deficits and decrease Aβ deposits in Tg2576 mice [176,177,178,179] and APP23 mice [180], which is consistent with clinical trial results [181,182,183,184]. Therefore, APP-Tg mice should be useful tools for developing new therapeutics aimed at suppressing the progression of AD. Drug repositioning is an attractive method for finding novel therapeutics. The nonsteroidal anti-inflammatory drug ibuprofen reduces amyloid deposits and inflammation [185] and increases memory and synaptic plasticity in Tg2576 mice [186] and 3×Tg-AD mice [187], but these effects are not seen in 5XFAD mice [188]. However, ibuprofen has both positive [189] and negative effects [190] in clinical trials. The peroxisome proliferator-activated receptor γ agonists rosiglitazone and pioglitazone also ameliorate hippocampus-dependent memory impairment in Tg2576 mice [73,191], APP/PS1 mice [192], and 3×Tg-AD mice [193], but the results from clinical trials were not consistent with those from the animal experiments [194,195,196,197]. Multiple AD clinical trials using statin drugs, which decrease blood cholesterol, were performed. However, a phase IV clinical trial with the statin simvastatin was unsuccessful [198]. As a result, further studies with statins are needed to understand their mechanism of action in AD treatment. Rifampicin also reduces Aβ accumulation in Tg2576 mice [199], but little is known about the mechanism by which it modulates AD-related amyloid deposits.
Several promising novel therapeutic candidates for the treatment of AD have emerged. BACE and γ-secretase inhibitors suppress the production of insoluble Aβ, reduce Aβ deposition, and ameliorate cognitive impairments in Tg2576 [200], PDAPP [201], and APPswe/PSEN1dE9 mice [202]. However, because these inhibitors aggravate cognitive impairments and have side effects, no inhibitors passed preclinical trials [203,204,205]. The usefulness of Aβ42 vaccination was first observed in PDAPP mice [206]. PDAPP mice that received this vaccine showed reduced Aβ depositions. The vaccination ameliorated memory abnormalities in APP+PS1 [207] and TgCRND8 mice [96]. The vaccination was also used to develop anti-Aβ antibodies, and their administration induced a dramatic reduction in Aβ and recovery of memory deficits in Tg2576 [208,209] and PDAPP [210,211] mice without side effects. The anti-Aβ antibodies were administrated in clinical trials, but patients developed meningoencephalitis, and the trial was discontinued [212]. However, a follow-up study with AD patients from this trial who experienced negative side effects shows that patients have long-term clinical benefits such as amelioration in cognitive impairment and a reduction in CSF tau levels [213]. This led to the development of humanized anti-Aβ antibody drugs [214,215,216,217] and Aβ vaccines [218], and several clinical trials with anti-human Aβ monoclonal antibodies were performed [219,220]. Unfortunately, most of the trials were either unsuccessful or did not show a dramatic improvement in AD pathology [219,220]. Given that the amyloid cascade hypothesis states that Aβ is first deposited extracellularly 20–30 years before the onset of cognitive and memory deficits in the FAD brain, immunotherapy trials may have been started too late in the clinical phase, when too much Aβ had accumulated, and the Aβ cascade was already irrevocably initiated [219]. Despite these controversies, development of these antibody drugs is ongoing under a conventional clinical trial. Therefore, it is expected that new trials will be initiated earlier in the course of AD and that these antibodies will play a central role in future treatment strategies [219]. Hyperphosphorylated tau vaccines were also developed using AD model mice that overexpress mutant human tau, but their effects have yet to be confirmed in clinical trials [221,222]. Given the challenges and a large number of failures, development of new therapeutic drugs for AD treatment is strongly desired.

8. Conclusions

We introduced the molecular mechanisms of AD, including the amyloid cascade and tau propagation hypotheses, and biomarkers of AD. We also described representative APP-Tg mice in brief. Moreover, we indicated whether APP-Tg mice are appropriate as an AD animal model, concluding that they should be a partial model of AD because the phenotypes of these mice differ from those of AD patients. Although AD has been studied for many years, no effective therapeutic strategies for AD have been found. In terms of translational research, the gap between basic and clinical drug discovery screening methods may be one of the reasons why candidate drugs do not easily become therapeutic drugs for AD. In order to elucidate the mechanisms of AD onset, therapeutic strategies that initiate from a new point of view or unconventional methods are desired. Namely, new animal models that faithfully recapitulate AD pathology and new technologies that accurately detect AD symptoms are required. To solve some of these problems, new diagnostics and therapeutic approaches, such as APP-KI mice, touchscreen-associated testing techniques, and anti-Aβ drugs, were developed. As the development of these novel approaches increases, it is hoped that AD treatment targets that reduce or eliminate Aβ and NFT deposits will emerge. In addition, the accumulating studies that focus on the more common AD treatment approaches will be useful for supporting the detection of biochemical parameters and accurately monitoring AD progression.

Author Contributions

Conceptualization, T.N. and H.M.; writing—original draft preparation, T.N.; writing—review and editing, T.N. and H.M.; supervision, K.Y. and H.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was partially supported by the following funding sources: Grants-in-Aid for Scientific Research (19H03532, 19H05017, 19K21811, 20H01037) from MEXT; a grant for biomedical research from the SRF, Japan; a grant from the Takeda Science Foundation; a grant from the Mishima Kaiun Memorial Foundation; the Hori Sciences and Arts Foundation; and a grant from the Strategic Research Program for Brain Science from the Japan Agency for Medical Research and Development (AMED).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Not applicable.

Conflicts of Interest

The authors declare no competing financial interests.

References

  1. World Health Organization. 10 Facts on Dementia 2019. 2019. Available online: https://www.who.int/features/factfiles/dementia/en/ (accessed on 29 April 2021).
  2. Alzheimer’s Association. 2020 Alzheimer’s disease facts and figures. Alzheimer’s Dement. 2020. [Google Scholar] [CrossRef]
  3. Querfurth, H.W.; LaFerla, F.M. Alzheimer’s disease. N. Engl. J. Med. 2010, 362, 329–344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Bateman, R.J.; Xiong, C.; Benzinger, T.L.; Fagan, A.M.; Goate, A.; Fox, N.C.; Marcus, D.S.; Cairns, N.J.; Xie, X.; Blazey, T.M.; et al. Clinical and biomarker changes in dominantly inherited Alzheimer’s disease. N. Engl. J. Med. 2012, 367, 795–804. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. McKhann, G.M.; Knopman, D.S.; Chertkow, H.; Hyman, B.T.; Jack, C.R., Jr.; Kawas, C.H.; Klunk, W.E.; Koroshetz, W.J.; Manly, J.J.; Mayeux, R.; et al. The diagnosis of dementia due to Alzheimer’s disease: Recommendations from the National Institute on Aging-Alzheimer’s Association workgroups on diagnostic guidelines for Alzheimer’s disease. Alzheimer’s Dement. 2011, 7, 263–269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Vermunt, L.; Sikkes, S.A.M.; van den Hout, A.; Handels, R.; Bos, I.; van der Flier, W.M.; Kern, S.; Ousset, P.J.; Maruff, P.; Skoog, I.; et al. Duration of preclinical, prodromal, and dementia stages of Alzheimer’s disease in relation to age, sex, and APOE genotype. Alzheimer’s Dement. 2019, 15, 888–898. [Google Scholar] [CrossRef] [PubMed]
  7. Kim, D.H.; Jang, Y.S.; Jeon, W.K.; Han, J.S. Assessment of Cognitive Phenotyping in Inbred, Genetically Modified Mice, and Transgenic Mouse Models of Alzheimer’s Disease. Exp. Neurobiol. 2019, 28, 146–157. [Google Scholar] [CrossRef] [PubMed]
  8. Vyas, Y.; Montgomery, J.M.; Cheyne, J.E. Hippocampal Deficits in Amyloid-β-Related Rodent Models of Alzheimer’s Disease. Front. Neurosci. 2020, 14, 266. [Google Scholar] [CrossRef]
  9. Hardy, J.; Allsop, D. Amyloid deposition as the central event in the aetiology of Alzheimer’s disease. Trends Pharmacol. Sci. 1991, 12, 383–388. [Google Scholar] [CrossRef]
  10. Glenner, G.G.; Wong, C.W. Alzheimer’s disease: Initial report of the purification and characterization of a novel cerebrovascular amyloid protein. Biochem. Biophys. Res. Commun. 1984, 120, 885–890. [Google Scholar] [CrossRef]
  11. Robakis, N.K.; Ramakrishna, N.; Wolfe, G.; Wisniewski, H.M. Molecular cloning and characterization of a cDNA encoding the cerebrovascular and the neuritic plaque amyloid peptides. Proc. Natl. Acad. Sci. USA 1987, 84, 4190–4194. [Google Scholar] [CrossRef] [Green Version]
  12. Tanzi, R.E.; Gusella, J.F.; Watkins, P.C.; Bruns, G.A.; St George-Hyslop, P.; Van Keuren, M.L.; Patterson, D.; Pagan, S.; Kurnit, D.M.; Neve, R.L. Amyloid beta protein gene: cDNA, mRNA distribution, and genetic linkage near the Alzheimer locus. Science 1987, 235, 880–884. [Google Scholar] [CrossRef] [PubMed]
  13. Kang, J.; Lemaire, H.G.; Unterbeck, A.; Salbaum, J.M.; Masters, C.L.; Grzeschik, K.H.; Multhaup, G.; Beyreuther, K.; Müller-Hill, B. The precursor of Alzheimer’s disease amyloid A4 protein resembles a cell-surface receptor. Nature 1987, 325, 733–736. [Google Scholar] [CrossRef] [PubMed]
  14. Haass, C.; Kaether, C.; Thinakaran, G.; Sisodia, S. Trafficking and proteolytic processing of APP. Cold Spring Harb. Perspect. Med. 2012, 2, a006270. [Google Scholar] [CrossRef] [PubMed]
  15. Allinson, T.M.; Parkin, E.T.; Turner, A.J.; Hooper, N.M. ADAMs family members as amyloid precursor protein alpha-secretases. J. Neurosci. Res. 2003, 74, 342–352. [Google Scholar] [CrossRef]
  16. Wang, J.; Dickson, D.W.; Trojanowski, J.Q.; Lee, V.M. The levels of soluble versus insoluble brain Abeta distinguish Alzheimer’s disease from normal and pathologic aging. Exp. Neurol. 1999, 158, 328–337. [Google Scholar] [CrossRef]
  17. Yankner, B.A.; Duffy, L.K.; Kirschner, D.A. Neurotrophic and neurotoxic effects of amyloid beta protein: Reversal by tachykinin neuropeptides. Science 1990, 250, 279–282. [Google Scholar] [CrossRef]
  18. Tomlinson, B.E.; Irving, D.; Blessed, G. Cell loss in the locus coeruleus in senile dementia of Alzheimer type. J. Neurol. Sci. 1981, 49, 419–428. [Google Scholar] [CrossRef]
  19. Whitehouse, P.J.; Price, D.L.; Clark, A.W.; Coyle, J.T.; DeLong, M.R. Alzheimer disease: Evidence for selective loss of cholinergic neurons in the nucleus basalis. Ann. Neurol. 1981, 10, 122–126. [Google Scholar] [CrossRef]
  20. Mullan, M.; Crawford, F.; Axelman, K.; Houlden, H.; Lilius, L.; Winblad, B.; Lannfelt, L. A pathogenic mutation for probable Alzheimer’s disease in the APP gene at the N-terminus of beta-amyloid. Nat. Genet. 1992, 1, 345–347. [Google Scholar] [CrossRef]
  21. Di Fede, G.; Catania, M.; Morbin, M.; Rossi, G.; Suardi, S.; Mazzoleni, G.; Merlin, M.; Giovagnoli, A.R.; Prioni, S.; Erbetta, A.; et al. A recessive mutation in the APP gene with dominant-negative effect on amyloidogenesis. Science 2009, 323, 1473–1477. [Google Scholar] [CrossRef] [Green Version]
  22. Zhang, S.; Wang, Z.; Cai, F.; Zhang, M.; Wu, Y.; Zhang, J.; Song, W. BACE1 Cleavage Site Selection Critical for Amyloidogenesis and Alzheimer’s Pathogenesis. J. Neurosci. 2017, 37, 6915–6925. [Google Scholar] [CrossRef] [Green Version]
  23. Sisodia, S.S.; Koo, E.H.; Beyreuther, K.; Unterbeck, A.; Price, D.L. Evidence that beta-amyloid protein in Alzheimer’s disease is not derived by normal processing. Science 1990, 248, 492–495. [Google Scholar] [CrossRef] [PubMed]
  24. Ahmed, M.; Davis, J.; Aucoin, D.; Sato, T.; Ahuja, S.; Aimoto, S.; Elliott, J.I.; Van Nostrand, W.E.; Smith, S.O. Structural conversion of neurotoxic amyloid-beta(1-42) oligomers to fibrils. Nat. Struct. Mol. Biol. 2010, 17, 561–567. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Mayeux, R.; Tang, M.X.; Jacobs, D.M.; Manly, J.; Bell, K.; Merchant, C.; Small, S.A.; Stern, Y.; Wisniewski, H.M.; Mehta, P.D. Plasma amyloid beta-peptide 1-42 and incipient Alzheimer’s disease. Ann. Neurol. 1999, 46, 412–416. [Google Scholar] [CrossRef]
  26. Frost, B.; Jacks, R.L.; Diamond, M.I. Propagation of tau misfolding from the outside to the inside of a cell. J. Biol. Chem. 2009, 284, 12845–12852. [Google Scholar] [CrossRef] [Green Version]
  27. Weingarten, M.D.; Lockwood, A.H.; Hwo, S.Y.; Kirschner, M.W. A protein factor essential for microtubule assembly. Proc. Natl. Acad. Sci. USA 1975, 72, 1858–1862. [Google Scholar] [CrossRef] [Green Version]
  28. Goedert, M.; Wischik, C.M.; Crowther, R.A.; Walker, J.E.; Klug, A. Cloning and sequencing of the cDNA encoding a core protein of the paired helical filament of Alzheimer disease: Identification as the microtubule-associated protein tau. Proc. Natl. Acad. Sci. USA 1988, 85, 4051–4055. [Google Scholar] [CrossRef] [Green Version]
  29. Cleveland, D.W.; Hwo, S.Y.; Kirschner, M.W. Purification of tau, a microtubule-associated protein that induces assembly of microtubules from purified tubulin. J. Mol. Biol. 1977, 116, 207–225. [Google Scholar] [CrossRef]
  30. Goedert, M.; Spillantini, M.G.; Jakes, R.; Rutherford, D.; Crowther, R.A. Multiple isoforms of human microtubule-associated protein tau: Sequences and localization in neurofibrillary tangles of Alzheimer’s disease. Neuron 1989, 3, 519–526. [Google Scholar] [CrossRef]
  31. Adams, S.J.; DeTure, M.A.; McBride, M.; Dickson, D.W.; Petrucelli, L. Three repeat isoforms of tau inhibit assembly of four repeat tau filaments. PLoS ONE 2010, 5, e10810. [Google Scholar] [CrossRef]
  32. Harada, A.; Oguchi, K.; Okabe, S.; Kuno, J.; Terada, S.; Ohshima, T.; Sato-Yoshitake, R.; Takei, Y.; Noda, T.; Hirokawa, N. Altered microtubule organization in small-calibre axons of mice lacking tau protein. Nature 1994, 369, 488–491. [Google Scholar] [CrossRef] [PubMed]
  33. Noble, W.; Hanger, D.P.; Miller, C.C.; Lovestone, S. The importance of tau phosphorylation for neurodegenerative diseases. Front. Neurol. 2013, 4, 83. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Hanger, D.P.; Byers, H.L.; Wray, S.; Leung, K.Y.; Saxton, M.J.; Seereeram, A.; Reynolds, C.H.; Ward, M.A.; Anderton, B.H. Novel phosphorylation sites in tau from Alzheimer brain support a role for casein kinase 1 in disease pathogenesis. J. Biol. Chem. 2007, 282, 23645–23654. [Google Scholar] [CrossRef] [Green Version]
  35. Tavares, I.A.; Touma, D.; Lynham, S.; Troakes, C.; Schober, M.; Causevic, M.; Garg, R.; Noble, W.; Killick, R.; Bodi, I.; et al. Prostate-derived sterile 20-like kinases (PSKs/TAOKs) phosphorylate tau protein and are activated in tangle-bearing neurons in Alzheimer disease. J. Biol. Chem. 2013, 288, 15418–15429. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Tamaoka, A.; Sawamura, N.; Fukushima, T.; Shoji, S.; Matsubara, E.; Shoji, M.; Hirai, S.; Furiya, Y.; Endoh, R.; Mori, H. Amyloid beta protein 42(43) in cerebrospinal fluid of patients with Alzheimer’s disease. J. Neurol. Sci. 1997, 148, 41–45. [Google Scholar] [CrossRef]
  37. Skoog, I.; Davidsson, P.; Aevarsson, O.; Vanderstichele, H.; Vanmechelen, E.; Blennow, K. Cerebrospinal fluid beta-amyloid 42 is reduced before the onset of sporadic dementia: A population-based study in 85-year-olds. Dement. Geriatr. Cogn. Disord. 2003, 15, 169–176. [Google Scholar] [CrossRef]
  38. Vandermeeren, M.; Mercken, M.; Vanmechelen, E.; Six, J.; van de Voorde, A.; Martin, J.J.; Cras, P. Detection of tau proteins in normal and Alzheimer’s disease cerebrospinal fluid with a sensitive sandwich enzyme-linked immunosorbent assay. J. Neurochem. 1993, 61, 1828–1834. [Google Scholar] [CrossRef] [PubMed]
  39. Arai, H.; Terajima, M.; Miura, M.; Higuchi, S.; Muramatsu, T.; Machida, N.; Seiki, H.; Takase, S.; Clark, C.M.; Lee, V.M.; et al. Tau in cerebrospinal fluid: A potential diagnostic marker in Alzheimer’s disease. Ann. Neurol. 1995, 38, 649–652. [Google Scholar] [CrossRef]
  40. Motter, R.; Vigo-Pelfrey, C.; Kholodenko, D.; Barbour, R.; Johnson-Wood, K.; Galasko, D.; Chang, L.; Miller, B.; Clark, C.; Green, R.; et al. Reduction of beta-amyloid peptide42 in the cerebrospinal fluid of patients with Alzheimer’s disease. Ann. Neurol. 1995, 38, 643–648. [Google Scholar] [CrossRef]
  41. Shoji, M.; Matsubara, E.; Kanai, M.; Watanabe, M.; Nakamura, T.; Tomidokoro, Y.; Shizuka, M.; Wakabayashi, K.; Igeta, Y.; Ikeda, Y.; et al. Combination assay of CSF tau, A beta 1-40 and A beta 1-42(43) as a biochemical marker of Alzheimer’s disease. J. Neurol. Sci. 1998, 158, 134–140. [Google Scholar] [CrossRef]
  42. Fukumoto, H.; Tokuda, T.; Kasai, T.; Ishigami, N.; Hidaka, H.; Kondo, M.; Allsop, D.; Nakagawa, M. High-molecular-weight beta-amyloid oligomers are elevated in cerebrospinal fluid of Alzheimer patients. FASEB J. 2010, 24, 2716–2726. [Google Scholar] [CrossRef] [PubMed]
  43. Murakami, K.; Tokuda, M.; Suzuki, T.; Irie, Y.; Hanaki, M.; Izuo, N.; Monobe, Y.; Akagi, K.; Ishii, R.; Tatebe, H.; et al. Monoclonal antibody with conformational specificity for a toxic conformer of amyloid β42 and its application toward the Alzheimer’s disease diagnosis. Sci. Rep. 2016, 6, 29038. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Klunk, W.E.; Engler, H.; Nordberg, A.; Wang, Y.; Blomqvist, G.; Holt, D.P.; Bergström, M.; Savitcheva, I.; Huang, G.F.; Estrada, S.; et al. Imaging brain amyloid in Alzheimer’s disease with Pittsburgh Compound-B. Ann. Neurol. 2004, 55, 306–319. [Google Scholar] [CrossRef]
  45. Villemagne, V.L.; Burnham, S.; Bourgeat, P.; Brown, B.; Ellis, K.A.; Salvado, O.; Szoeke, C.; Macaulay, S.L.; Martins, R.; Maruff, P.; et al. Amyloid β deposition, neurodegeneration, and cognitive decline in sporadic Alzheimer’s disease: A prospective cohort study. Lancet Neurol. 2013, 12, 357–367. [Google Scholar] [CrossRef]
  46. Yang, L.; Rieves, D.; Ganley, C. Brain amyloid imaging--FDA approval of florbetapir F18 injection. N. Engl. J. Med. 2012, 367, 885–887. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Lister-James, J.; Pontecorvo, M.J.; Clark, C.; Joshi, A.D.; Mintun, M.A.; Zhang, W.; Lim, N.; Zhuang, Z.; Golding, G.; Choi, S.R.; et al. Florbetapir f-18: A histopathologically validated Beta-amyloid positron emission tomography imaging agent. Semin. Nucl. Med. 2011, 41, 300–304. [Google Scholar] [CrossRef]
  48. Harada, R.; Okamura, N.; Furumoto, S.; Furukawa, K.; Ishiki, A.; Tomita, N.; Tago, T.; Hiraoka, K.; Watanuki, S.; Shidahara, M.; et al. 18F-THK5351: A Novel PET Radiotracer for Imaging Neurofibrillary Pathology in Alzheimer Disease. J. Nucl. Med. 2016, 57, 208–214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Johnson, K.A.; Schultz, A.; Betensky, R.A.; Becker, J.A.; Sepulcre, J.; Rentz, D.; Mormino, E.; Chhatwal, J.; Amariglio, R.; Papp, K.; et al. Tau positron emission tomographic imaging in aging and early Alzheimer disease. Ann. Neurol. 2016, 79, 110–119. [Google Scholar] [CrossRef] [Green Version]
  50. Cagnin, A.; Brooks, D.J.; Kennedy, A.M.; Gunn, R.N.; Myers, R.; Turkheimer, F.E.; Jones, T.; Banati, R.B. In-vivo measurement of activated microglia in dementia. Lancet 2001, 358, 461–467. [Google Scholar] [CrossRef]
  51. Yasuno, F.; Ota, M.; Kosaka, J.; Ito, H.; Higuchi, M.; Doronbekov, T.K.; Nozaki, S.; Fujimura, Y.; Koeda, M.; Asada, T.; et al. Increased binding of peripheral benzodiazepine receptor in Alzheimer’s disease measured by positron emission tomography with [11C]DAA1106. Biol. Psychiatry 2008, 64, 835–841. [Google Scholar] [CrossRef]
  52. Rembach, A.; Faux, N.G.; Watt, A.D.; Pertile, K.K.; Rumble, R.L.; Trounson, B.O.; Fowler, C.J.; Roberts, B.R.; Perez, K.A.; Li, Q.X.; et al. Changes in plasma amyloid beta in a longitudinal study of aging and Alzheimer’s disease. Alzheimer’s Dement. 2014, 10, 53–61. [Google Scholar] [CrossRef] [PubMed]
  53. Wood, H. Alzheimer disease: Biomarkers of AD risk - the end of the road for plasma amyloid-β? Nat. Rev. Neurol. 2016, 12, 613. [Google Scholar] [CrossRef] [PubMed]
  54. Lövheim, H.; Elgh, F.; Johansson, A.; Zetterberg, H.; Blennow, K.; Hallmans, G.; Eriksson, S. Plasma concentrations of free amyloid β cannot predict the development of Alzheimer’s disease. Alzheimer’s Dement. 2017, 13, 778–782. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Kaneko, N.; Nakamura, A.; Washimi, Y.; Kato, T.; Sakurai, T.; Arahata, Y.; Bundo, M.; Takeda, A.; Niida, S.; Ito, K.; et al. Novel plasma biomarker surrogating cerebral amyloid deposition. Proc. Jpn. Acad. Ser. B Phys. Biol. Sci. 2014, 90, 353–364. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Ovod, V.; Ramsey, K.N.; Mawuenyega, K.G.; Bollinger, J.G.; Hicks, T.; Schneider, T.; Sullivan, M.; Paumier, K.; Holtzman, D.M.; Morris, J.C.; et al. Amyloid β concentrations and stable isotope labeling kinetics of human plasma specific to central nervous system amyloidosis. Alzheimer’s Dement. 2017, 13, 841–849. [Google Scholar] [CrossRef] [PubMed]
  57. Nakamura, A.; Kaneko, N.; Villemagne, V.L.; Kato, T.; Doecke, J.; Doré, V.; Fowler, C.; Li, Q.X.; Martins, R.; Rowe, C.; et al. High performance plasma amyloid-β biomarkers for Alzheimer’s disease. Nature 2018, 554, 249–254. [Google Scholar] [CrossRef]
  58. Fandos, N.; Pérez-Grijalba, V.; Pesini, P.; Olmos, S.; Bossa, M.; Villemagne, V.L.; Doecke, J.; Fowler, C.; Masters, C.L.; Sarasa, M. Plasma amyloid β 42/40 ratios as biomarkers for amyloid β cerebral deposition in cognitively normal individuals. Alzheimers Dement (Amst). 2017, 8, 179–187. [Google Scholar] [CrossRef]
  59. Nabers, A.; Perna, L.; Lange, J.; Mons, U.; Schartner, J.; Güldenhaupt, J.; Saum, K.U.; Janelidze, S.; Holleczek, B.; Rujescu, D.; et al. Amyloid blood biomarker detects Alzheimer’s disease. EMBO Mol. Med. 2018, 10. [Google Scholar] [CrossRef]
  60. Tatebe, H.; Kasai, T.; Ohmichi, T.; Kishi, Y.; Kakeya, T.; Waragai, M.; Kondo, M.; Allsop, D.; Tokuda, T. Quantification of plasma phosphorylated tau to use as a biomarker for brain Alzheimer pathology: Pilot case-control studies including patients with Alzheimer’s disease and down syndrome. Mol. Neurodegener. 2017, 12, 63. [Google Scholar] [CrossRef] [Green Version]
  61. Janelidze, S.; Mattsson, N.; Palmqvist, S.; Smith, R.; Beach, T.G.; Serrano, G.E.; Chai, X.; Proctor, N.K.; Eichenlaub, U.; Zetterberg, H.; et al. Plasma P-tau181 in Alzheimer’s disease: Relationship to other biomarkers, differential diagnosis, neuropathology and longitudinal progression to Alzheimer’s dementia. Nat. Med. 2020, 26, 379–386. [Google Scholar] [CrossRef]
  62. Thijssen, E.H.; La Joie, R.; Wolf, A.; Strom, A.; Wang, P.; Iaccarino, L.; Bourakova, V.; Cobigo, Y.; Heuer, H.; Spina, S.; et al. Diagnostic value of plasma phosphorylated tau181 in Alzheimer’s disease and frontotemporal lobar degeneration. Nat. Med. 2020, 26, 387–397. [Google Scholar] [CrossRef] [PubMed]
  63. Bettcher, B.M.; Johnson, S.C.; Fitch, R.; Casaletto, K.B.; Heffernan, K.S.; Asthana, S.; Zetterberg, H.; Blennow, K.; Carlsson, C.M.; Neuhaus, J.; et al. Cerebrospinal Fluid and Plasma Levels of Inflammation Differentially Relate to CNS Markers of Alzheimer’s Disease Pathology and Neuronal Damage. J. Alzheimers Dis. 2018, 62, 385–397. [Google Scholar] [CrossRef] [PubMed]
  64. Mattsson, N.; Andreasson, U.; Zetterberg, H.; Blennow, K. Association of Plasma Neurofilament Light With Neurodegeneration in Patients With Alzheimer Disease. JAMA Neurol. 2017, 74, 557–566. [Google Scholar] [CrossRef] [PubMed]
  65. Shigemizu, D.; Akiyama, S.; Asanomi, Y.; Boroevich, K.A.; Sharma, A.; Tsunoda, T.; Matsukuma, K.; Ichikawa, M.; Sudo, H.; Takizawa, S.; et al. Risk prediction models for dementia constructed by supervised principal component analysis using miRNA expression data. Commun Biol. 2019, 2, 77. [Google Scholar] [CrossRef] [Green Version]
  66. Wennberg, A.M.V.; Hagen, C.E.; Machulda, M.M.; Hollman, J.H.; Roberts, R.O.; Knopman, D.S.; Petersen, R.C.; Mielke, M.M. The association between peripheral total IGF-1, IGFBP-3, and IGF-1/IGFBP-3 and functional and cognitive outcomes in the Mayo Clinic Study of Aging. Neurobiol. Aging 2018, 66, 68–74. [Google Scholar] [CrossRef]
  67. Liu, S.; Suzuki, H.; Ito, H.; Korenaga, T.; Akatsu, H.; Meno, K.; Uchida, K. Serum levels of proteins involved in amyloid-β clearance are related to cognitive decline and neuroimaging changes in mild cognitive impairment. Alzheimers Dement (Amst). 2019, 11, 85–97. [Google Scholar] [CrossRef]
  68. Cruts, M.; van Duijn, C.M.; Backhovens, H.; Van den Broeck, M.; Wehnert, A.; Serneels, S.; Sherrington, R.; Hutton, M.; Hardy, J.; St George-Hyslop, P.H.; et al. Estimation of the genetic contribution of presenilin-1 and -2 mutations in a population-based study of presenile Alzheimer disease. Hum. Mol. Genet. 1998, 7, 43–51. [Google Scholar] [CrossRef] [PubMed]
  69. Cuyvers, E.; Sleegers, K. Genetic variations underlying Alzheimer’s disease: Evidence from genome-wide association studies and beyond. Lancet Neurol. 2016, 15, 857–868. [Google Scholar] [CrossRef]
  70. Hsiao, K.; Chapman, P.; Nilsen, S.; Eckman, C.; Harigaya, Y.; Younkin, S.; Yang, F.; Cole, G. Correlative memory deficits, Abeta elevation, and amyloid plaques in transgenic mice. Science 1996, 274, 99–102. [Google Scholar] [CrossRef]
  71. Benzing, W.C.; Wujek, J.R.; Ward, E.K.; Shaffer, D.; Ashe, K.H.; Younkin, S.G.; Brunden, K.R. Evidence for glial-mediated inflammation in aged APP(SW) transgenic mice. Neurobiol. Aging 1999, 20, 581–589. [Google Scholar] [CrossRef]
  72. Corcoran, K.A.; Lu, Y.; Turner, R.S.; Maren, S. Overexpression of hAPPswe impairs rewarded alternation and contextual fear conditioning in a transgenic mouse model of Alzheimer’s disease. Learn. Mem. 2002, 9, 243–252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Cortez, I.; Hernandez, C.M.; Dineley, K.T. Enhancement of select cognitive domains with rosiglitazone implicates dorsal hippocampus circuitry sensitive to PPARγ agonism in an Alzheimer’s mouse model. Brain Behav. 2021, 11, e01973. [Google Scholar] [CrossRef] [PubMed]
  74. Barnes, P.; Good, M. Impaired Pavlovian cued fear conditioning in Tg2576 mice expressing a human mutant amyloid precursor protein gene. Behav. Brain Res. 2005, 157, 107–117. [Google Scholar] [CrossRef] [PubMed]
  75. Scopa, C.; Marrocco, F.; Latina, V.; Ruggeri, F.; Corvaglia, V.; La Regina, F.; Ammassari-Teule, M.; Middei, S.; Amadoro, G.; Meli, G.; et al. Impaired adult neurogenesis is an early event in Alzheimer’s disease neurodegeneration, mediated by intracellular Aβ oligomers. Cell Death Differ. 2020, 27, 934–948. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Braak, H.; Braak, E. Neuropathological stageing of Alzheimer-related changes. Acta Neuropathol. 1991, 82, 239–259. [Google Scholar] [CrossRef] [PubMed]
  77. Yang, F.; Uéda, K.; Chen, P.; Ashe, K.H.; Cole, G.M. Plaque-associated alpha-synuclein (NACP) pathology in aged transgenic mice expressing amyloid precursor protein. Brain Res. 2000, 853, 381–383. [Google Scholar] [CrossRef]
  78. Sturchler-Pierrat, C.; Abramowski, D.; Duke, M.; Wiederhold, K.H.; Mistl, C.; Rothacher, S.; Ledermann, B.; Bürki, K.; Frey, P.; Paganetti, P.A.; et al. Two amyloid precursor protein transgenic mouse models with Alzheimer disease-like pathology. Proc. Natl. Acad. Sci. USA 1997, 94, 13287–13292. [Google Scholar] [CrossRef] [Green Version]
  79. Calhoun, M.E.; Wiederhold, K.H.; Abramowski, D.; Phinney, A.L.; Probst, A.; Sturchler-Pierrat, C.; Staufenbiel, M.; Sommer, B.; Jucker, M. Neuron loss in APP transgenic mice. Nature 1998, 395, 755–756. [Google Scholar] [CrossRef]
  80. Bondolfi, L.; Calhoun, M.; Ermini, F.; Kuhn, H.G.; Wiederhold, K.H.; Walker, L.; Staufenbiel, M.; Jucker, M. Amyloid-associated neuron loss and gliogenesis in the neocortex of amyloid precursor protein transgenic mice. J. Neurosci. 2002, 22, 515–522. [Google Scholar] [CrossRef] [Green Version]
  81. Calhoun, M.E.; Burgermeister, P.; Phinney, A.L.; Stalder, M.; Tolnay, M.; Wiederhold, K.H.; Abramowski, D.; Sturchler-Pierrat, C.; Sommer, B.; Staufenbiel, M.; et al. Neuronal overexpression of mutant amyloid precursor protein results in prominent deposition of cerebrovascular amyloid. Proc. Natl. Acad. Sci. USA 1999, 96, 14088–14093. [Google Scholar] [CrossRef] [Green Version]
  82. Kelly, P.H.; Bondolfi, L.; Hunziker, D.; Schlecht, H.P.; Carver, K.; Maguire, E.; Abramowski, D.; Wiederhold, K.H.; Sturchler-Pierrat, C.; Jucker, M.; et al. Progressive age-related impairment of cognitive behavior in APP23 transgenic mice. Neurobiol. Aging 2003, 24, 365–378. [Google Scholar] [CrossRef]
  83. Van Dam, D.; D’Hooge, R.; Staufenbiel, M.; Van Ginneken, C.; Van Meir, F.; De Deyn, P.P. Age-dependent cognitive decline in the APP23 model precedes amyloid deposition. Eur. J. Neurosci. 2003, 17, 388–396. [Google Scholar] [CrossRef]
  84. Liu, X.; Yamashita, T.; Shang, J.; Shi, X.; Morihara, R.; Huang, Y.; Sato, K.; Takemoto, M.; Hishikawa, N.; Ohta, Y.; et al. Clinical and Pathological Benefit of Twendee X in Alzheimer’s Disease Transgenic Mice with Chronic Cerebral Hypoperfusion. J. Stroke Cerebrovasc. Dis. 2019, 28, 1993–2002. [Google Scholar] [CrossRef] [PubMed]
  85. Van Erum, J.; Van Dam, D.; Sheorajpanday, R.; De Deyn, P.P. Sleep architecture changes in the APP23 mouse model manifest at onset of cognitive deficits. Behav. Brain Res. 2019, 373, 112089. [Google Scholar] [CrossRef] [PubMed]
  86. Sorgdrager, F.; van Der Ley, C.P.; van Faassen, M.; Calus, E.; Nollen, E.A.; Kema, I.P.; van Dam, D.; De Deyn, P.P. The Effect of Tryptophan 2,3-Dioxygenase Inhibition on Kynurenine Metabolism and Cognitive Function in the APP23 Mouse Model of Alzheimer’s Disease. Int. J. Tryptophan Res. 2020, 13, 1178646920972657. [Google Scholar] [CrossRef] [PubMed]
  87. Chartier-Harlin, M.C.; Crawford, F.; Houlden, H.; Warren, A.; Hughes, D.; Fidani, L.; Goate, A.; Rossor, M.; Roques, P.; Hardy, J.; et al. Early-onset Alzheimer’s disease caused by mutations at codon 717 of the beta-amyloid precursor protein gene. Nature 1991, 353, 844–846. [Google Scholar] [CrossRef] [PubMed]
  88. Games, D.; Adams, D.; Alessandrini, R.; Barbour, R.; Berthelette, P.; Blackwell, C.; Carr, T.; Clemens, J.; Donaldson, T.; Gillespie, F.; et al. Alzheimer-type neuropathology in transgenic mice overexpressing V717F beta-amyloid precursor protein. Nature 1995, 373, 523–527. [Google Scholar] [CrossRef]
  89. Masliah, E.; Sisk, A.; Mallory, M.; Games, D. Neurofibrillary pathology in transgenic mice overexpressing V717F beta-amyloid precursor protein. J. Neuropathol. Exp. Neurol. 2001, 60, 357–368. [Google Scholar] [CrossRef] [Green Version]
  90. Dodart, J.C.; Meziane, H.; Mathis, C.; Bales, K.R.; Paul, S.M.; Ungerer, A. Behavioral disturbances in transgenic mice overexpressing the V717F beta-amyloid precursor protein. Behav. Neurosci. 1999, 113, 982–990. [Google Scholar] [CrossRef]
  91. Chen, G.; Chen, K.S.; Knox, J.; Inglis, J.; Bernard, A.; Martin, S.J.; Justice, A.; McConlogue, L.; Games, D.; Freedman, S.B.; et al. A learning deficit related to age and beta-amyloid plaques in a mouse model of Alzheimer’s disease. Nature 2000, 408, 975–979. [Google Scholar] [CrossRef]
  92. Hartman, R.E.; Izumi, Y.; Bales, K.R.; Paul, S.M.; Wozniak, D.F.; Holtzman, D.M. Treatment with an amyloid-beta antibody ameliorates plaque load, learning deficits, and hippocampal long-term potentiation in a mouse model of Alzheimer’s disease. J. Neurosci. 2005, 25, 6213–6220. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Chishti, M.A.; Yang, D.S.; Janus, C.; Phinney, A.L.; Horne, P.; Pearson, J.; Strome, R.; Zuker, N.; Loukides, J.; French, J.; et al. Early-onset amyloid deposition and cognitive deficits in transgenic mice expressing a double mutant form of amyloid precursor protein 695. J. Biol. Chem. 2001, 276, 21562–21570. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Bellucci, A.; Rosi, M.C.; Grossi, C.; Fiorentini, A.; Luccarini, I.; Casamenti, F. Abnormal processing of tau in the brain of aged TgCRND8 mice. Neurobiol. Dis. 2007, 27, 328–338. [Google Scholar] [CrossRef] [PubMed]
  95. Brautigam, H.; Steele, J.W.; Westaway, D.; Fraser, P.E.; St George-Hyslop, P.H.; Gandy, S.; Hof, P.R.; Dickstein, D.L. The isotropic fractionator provides evidence for differential loss of hippocampal neurons in two mouse models of Alzheimer’s disease. Mol. Neurodegener. 2012, 7, 58. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Janus, C.; Pearson, J.; McLaurin, J.; Mathews, P.M.; Jiang, Y.; Schmidt, S.D.; Chishti, M.A.; Horne, P.; Heslin, D.; French, J.; et al. A beta peptide immunization reduces behavioural impairment and plaques in a model of Alzheimer’s disease. Nature 2000, 408, 979–982. [Google Scholar] [CrossRef] [PubMed]
  97. Ambrée, O.; Touma, C.; Görtz, N.; Keyvani, K.; Paulus, W.; Palme, R.; Sachser, N. Activity changes and marked stereotypic behavior precede Abeta pathology in TgCRND8 Alzheimer mice. Neurobiol. Aging 2006, 27, 955–964. [Google Scholar] [CrossRef]
  98. Hamm, V.; Héraud, C.; Bott, J.B.; Herbeaux, K.; Strittmatter, C.; Mathis, C.; Goutagny, R. Differential contribution of APP metabolites to early cognitive deficits in a TgCRND8 mouse model of Alzheimer’s disease. Sci. Adv. 2017, 3, e1601068. [Google Scholar] [CrossRef] [Green Version]
  99. Xia, F.; Yiu, A.; Stone, S.S.D.; Oh, S.; Lozano, A.M.; Josselyn, S.A.; Frankland, P.W. Entorhinal Cortical Deep Brain Stimulation Rescues Memory Deficits in Both Young and Old Mice Genetically Engineered to Model Alzheimer’s Disease. Neuropsychopharmacology 2017, 42, 2493–2503. [Google Scholar] [CrossRef] [Green Version]
  100. Xian, Y.F.; Qu, C.; Liu, Y.; Ip, S.P.; Yuan, Q.J.; Yang, W.; Lin, Z.X. Magnolol Ameliorates Behavioral Impairments and Neuropathology in a Transgenic Mouse Model of Alzheimer’s Disease. Oxid. Med. Cell. Longev. 2020, 2020, 5920476. [Google Scholar] [CrossRef]
  101. Woodhouse, A.; Vickers, J.C.; Adlard, P.A.; Dickson, T.C. Dystrophic neurites in TgCRND8 and Tg2576 mice mimic human pathological brain aging. Neurobiol. Aging 2009, 30, 864–874. [Google Scholar] [CrossRef] [PubMed]
  102. Radde, R.; Bolmont, T.; Kaeser, S.A.; Coomaraswamy, J.; Lindau, D.; Stoltze, L.; Calhoun, M.E.; Jäggi, F.; Wolburg, H.; Gengler, S.; et al. Abeta42-driven cerebral amyloidosis in transgenic mice reveals early and robust pathology. EMBO Rep. 2006, 7, 940–946. [Google Scholar] [CrossRef] [Green Version]
  103. Rupp, N.J.; Wegenast-Braun, B.M.; Radde, R.; Calhoun, M.E.; Jucker, M. Early onset amyloid lesions lead to severe neuritic abnormalities and local, but not global neuron loss in APPPS1 transgenic mice. Neurobiol. Aging 2011, 32, 2324.e2321–2326. [Google Scholar] [CrossRef] [PubMed]
  104. Bittner, T.; Burgold, S.; Dorostkar, M.M.; Fuhrmann, M.; Wegenast-Braun, B.M.; Schmidt, B.; Kretzschmar, H.; Herms, J. Amyloid plaque formation precedes dendritic spine loss. Acta Neuropathol. 2012, 124, 797–807. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Serneels, L.; Van Biervliet, J.; Craessaerts, K.; Dejaegere, T.; Horré, K.; Van Houtvin, T.; Esselmann, H.; Paul, S.; Schäfer, M.K.; Berezovska, O.; et al. gamma-Secretase heterogeneity in the Aph1 subunit: Relevance for Alzheimer’s disease. Science 2009, 324, 639–642. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Montarolo, F.; Parolisi, R.; Hoxha, E.; Boda, E.; Tempia, F. Early enriched environment exposure protects spatial memory and accelerates amyloid plaque formation in APP(Swe)/PS1(L166P) mice. PLoS ONE 2013, 8, e69381. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Cifuentes, D.; Poittevin, M.; Dere, E.; Broquères-You, D.; Bonnin, P.; Benessiano, J.; Pocard, M.; Mariani, J.; Kubis, N.; Merkulova-Rainon, T.; et al. Hypertension accelerates the progression of Alzheimer-like pathology in a mouse model of the disease. Hypertension 2015, 65, 218–224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Wagner, L.K.; Gilling, K.E.; Schormann, E.; Kloetzel, P.M.; Heppner, F.L.; Krüger, E.; Prokop, S. Immunoproteasome deficiency alters microglial cytokine response and improves cognitive deficits in Alzheimer’s disease-like APPPS1 mice. Acta Neuropathol. Commun. 2017, 5, 52. [Google Scholar] [CrossRef]
  109. Van den Broeck, L.; Hansquine, P.; Callaerts-Vegh, Z.; D’Hooge, R. Impaired Reversal Learning in APPPS1-21 Mice in the Touchscreen Visual Discrimination Task. Front. Behav. Neurosci. 2019, 13, 92. [Google Scholar] [CrossRef] [Green Version]
  110. Oakley, H.; Cole, S.L.; Logan, S.; Maus, E.; Shao, P.; Craft, J.; Guillozet-Bongaarts, A.; Ohno, M.; Disterhoft, J.; Van Eldik, L.; et al. Intraneuronal beta-amyloid aggregates, neurodegeneration, and neuron loss in transgenic mice with five familial Alzheimer’s disease mutations: Potential factors in amyloid plaque formation. J. Neurosci. 2006, 26, 10129–10140. [Google Scholar] [CrossRef]
  111. Eimer, W.A.; Vassar, R. Neuron loss in the 5XFAD mouse model of Alzheimer’s disease correlates with intraneuronal Aβ42 accumulation and Caspase-3 activation. Mol. Neurodegener. 2013, 8, 2. [Google Scholar] [CrossRef] [Green Version]
  112. Chen, W.; Wang, M.; Zhu, M.; Xiong, W.; Qin, X.; Zhu, X. 14,15-Epoxyeicosatrienoic Acid Alleviates Pathology in a Mouse Model of Alzheimer’s Disease. J. Neurosci. 2020, 40, 8188–8203. [Google Scholar] [CrossRef] [PubMed]
  113. Ramasamy, V.S.; Samidurai, M.; Park, H.J.; Wang, M.; Park, R.Y.; Yu, S.Y.; Kang, H.K.; Hong, S.; Choi, W.S.; Lee, Y.Y.; et al. Avenanthramide-C Restores Impaired Plasticity and Cognition in Alzheimer’s Disease Model Mice. Mol. Neurobiol. 2020, 57, 315–330. [Google Scholar] [CrossRef] [PubMed]
  114. Ohno, M.; Chang, L.; Tseng, W.; Oakley, H.; Citron, M.; Klein, W.L.; Vassar, R.; Disterhoft, J.F. Temporal memory deficits in Alzheimer’s mouse models: Rescue by genetic deletion of BACE1. Eur. J. Neurosci. 2006, 23, 251–260. [Google Scholar] [CrossRef] [PubMed]
  115. Oh, S.B.; Kim, M.S.; Park, S.; Son, H.; Kim, S.Y.; Kim, M.S.; Jo, D.G.; Tak, E.; Lee, J.Y. Clusterin contributes to early stage of Alzheimer’s disease pathogenesis. Brain Pathol. 2019, 29, 217–231. [Google Scholar] [CrossRef] [PubMed]
  116. Zhong, L.; Xu, Y.; Zhuo, R.; Wang, T.; Wang, K.; Huang, R.; Wang, D.; Gao, Y.; Zhu, Y.; Sheng, X.; et al. Soluble TREM2 ameliorates pathological phenotypes by modulating microglial functions in an Alzheimer’s disease model. Nat. Commun. 2019, 10, 1365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Oddo, S.; Caccamo, A.; Shepherd, J.D.; Murphy, M.P.; Golde, T.E.; Kayed, R.; Metherate, R.; Mattson, M.P.; Akbari, Y.; LaFerla, F.M. Triple-transgenic model of Alzheimer’s disease with plaques and tangles: Intracellular Abeta and synaptic dysfunction. Neuron 2003, 39, 409–421. [Google Scholar] [CrossRef] [Green Version]
  118. Billings, L.M.; Oddo, S.; Green, K.N.; McGaugh, J.L.; LaFerla, F.M. Intraneuronal Abeta causes the onset of early Alzheimer’s disease-related cognitive deficits in transgenic mice. Neuron 2005, 45, 675–688. [Google Scholar] [CrossRef] [Green Version]
  119. Manaye, K.F.; Mouton, P.R.; Xu, G.; Drew, A.; Lei, D.L.; Sharma, Y.; Rebeck, G.W.; Turner, S. Age-related loss of noradrenergic neurons in the brains of triple transgenic mice. Age (Dordr). 2013, 35, 139–147. [Google Scholar] [CrossRef] [Green Version]
  120. Janelsins, M.C.; Mastrangelo, M.A.; Park, K.M.; Sudol, K.L.; Narrow, W.C.; Oddo, S.; LaFerla, F.M.; Callahan, L.M.; Federoff, H.J.; Bowers, W.J. Chronic neuron-specific tumor necrosis factor-alpha expression enhances the local inflammatory environment ultimately leading to neuronal death in 3xTg-AD mice. Am. J. Pathol. 2008, 173, 1768–1782. [Google Scholar] [CrossRef] [Green Version]
  121. Kastyak-Ibrahim, M.Z.; Di Curzio, D.L.; Buist, R.; Herrera, S.L.; Albensi, B.C.; Del Bigio, M.R.; Martin, M. Neurofibrillary tangles and plaques are not accompanied by white matter pathology in aged triple transgenic-Alzheimer disease mice. Magn. Reson. Imaging 2013, 31, 1515–1521. [Google Scholar] [CrossRef]
  122. Scuderi, C.; Bronzuoli, M.R.; Facchinetti, R.; Pace, L.; Ferraro, L.; Broad, K.D.; Serviddio, G.; Bellanti, F.; Palombelli, G.; Carpinelli, G.; et al. Ultramicronized palmitoylethanolamide rescues learning and memory impairments in a triple transgenic mouse model of Alzheimer’s disease by exerting anti-inflammatory and neuroprotective effects. Transl Psychiatry 2018, 8, 32. [Google Scholar] [CrossRef] [PubMed]
  123. Barone, E.; Tramutola, A.; Triani, F.; Calcagnini, S.; Di Domenico, F.; Ripoli, C.; Gaetani, S.; Grassi, C.; Butterfield, D.A.; Cassano, T.; et al. Biliverdin Reductase-A Mediates the Beneficial Effects of Intranasal Insulin in Alzheimer Disease. Mol. Neurobiol. 2019, 56, 2922–2943. [Google Scholar] [CrossRef]
  124. Escrig, A.; Canal, C.; Sanchis, P.; Fernández-Gayol, O.; Montilla, A.; Comes, G.; Molinero, A.; Giralt, M.; Giménez-Llort, L.; Becker-Pauly, C.; et al. IL-6 trans-signaling in the brain influences the behavioral and physio-pathological phenotype of the Tg2576 and 3xTgAD mouse models of Alzheimer’s disease. Brain. Behav. Immun. 2019, 82, 145–159. [Google Scholar] [CrossRef] [PubMed]
  125. Corsetti, V.; Borreca, A.; Latina, V.; Giacovazzo, G.; Pignataro, A.; Krashia, P.; Natale, F.; Cocco, S.; Rinaudo, M.; Malerba, F.; et al. Passive immunotherapy for N-truncated tau ameliorates the cognitive deficits in two mouse Alzheimer’s disease models. Brain Commun. 2020, 2, fcaa039. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Shen, Y.; Hua, L.; Yeh, C.K.; Shen, L.; Ying, M.; Zhang, Z.; Liu, G.; Li, S.; Chen, S.; Chen, X.; et al. Ultrasound with microbubbles improves memory, ameliorates pathology and modulates hippocampal proteomic changes in a triple transgenic mouse model of Alzheimer’s disease. Theranostics 2020, 10, 11794–11819. [Google Scholar] [CrossRef]
  127. Correia, S.C.; Machado, N.J.; Alves, M.G.; Oliveira, P.F.; Moreira, P.I. Intermittent Hypoxic Conditioning Rescues Cognition and Mitochondrial Bioenergetic Profile in the Triple Transgenic Mouse Model of Alzheimer’s Disease. Int. J. Mol. Sci. 2021, 22, 461. [Google Scholar] [CrossRef]
  128. Zannis, V.I.; Kardassis, D.; Zanni, E.E. Genetic mutations affecting human lipoproteins, their receptors, and their enzymes. Adv. Hum. Genet. 1993, 21, 145–319. [Google Scholar] [CrossRef]
  129. Liu, C.C.; Liu, C.C.; Kanekiyo, T.; Xu, H.; Bu, G. Apolipoprotein E and Alzheimer disease: Risk, mechanisms and therapy. Nat. Rev. Neurol. 2013, 9, 106–118. [Google Scholar] [CrossRef] [Green Version]
  130. Hoe, H.S.; Lee, K.J.; Carney, R.S.; Lee, J.; Markova, A.; Lee, J.Y.; Howell, B.W.; Hyman, B.T.; Pak, D.T.; Bu, G.; et al. Interaction of reelin with amyloid precursor protein promotes neurite outgrowth. J. Neurosci. 2009, 29, 7459–7473. [Google Scholar] [CrossRef]
  131. Bales, K.R.; Liu, F.; Wu, S.; Lin, S.; Koger, D.; DeLong, C.; Hansen, J.C.; Sullivan, P.M.; Paul, S.M. Human APOE isoform-dependent effects on brain beta-amyloid levels in PDAPP transgenic mice. J. Neurosci. 2009, 29, 6771–6779. [Google Scholar] [CrossRef]
  132. Bour, A.; Grootendorst, J.; Vogel, E.; Kelche, C.; Dodart, J.C.; Bales, K.; Moreau, P.H.; Sullivan, P.M.; Mathis, C. Middle-aged human apoE4 targeted-replacement mice show retention deficits on a wide range of spatial memory tasks. Behav. Brain Res. 2008, 193, 174–182. [Google Scholar] [CrossRef] [PubMed]
  133. Dumanis, S.B.; Tesoriero, J.A.; Babus, L.W.; Nguyen, M.T.; Trotter, J.H.; Ladu, M.J.; Weeber, E.J.; Turner, R.S.; Xu, B.; Rebeck, G.W.; et al. ApoE4 decreases spine density and dendritic complexity in cortical neurons in vivo. J. Neurosci. 2009, 29, 15317–15322. [Google Scholar] [CrossRef] [PubMed]
  134. Sullivan, P.M.; Han, B.; Liu, F.; Mace, B.E.; Ervin, J.F.; Wu, S.; Koger, D.; Paul, S.; Bales, K.R. Reduced levels of human apoE4 protein in an animal model of cognitive impairment. Neurobiol. Aging 2011, 32, 791–801. [Google Scholar] [CrossRef]
  135. Bales, K.R.; Verina, T.; Dodel, R.C.; Du, Y.; Altstiel, L.; Bender, M.; Hyslop, P.; Johnstone, E.M.; Little, S.P.; Cummins, D.J.; et al. Lack of apolipoprotein E dramatically reduces amyloid beta-peptide deposition. Nat. Genet. 1997, 17, 263–264. [Google Scholar] [CrossRef] [PubMed]
  136. Fryer, J.D.; Simmons, K.; Parsadanian, M.; Bales, K.R.; Paul, S.M.; Sullivan, P.M.; Holtzman, D.M. Human apolipoprotein E4 alters the amyloid-beta 40:42 ratio and promotes the formation of cerebral amyloid angiopathy in an amyloid precursor protein transgenic model. J. Neurosci. 2005, 25, 2803–2810. [Google Scholar] [CrossRef] [Green Version]
  137. Liao, F.; Zhang, T.J.; Jiang, H.; Lefton, K.B.; Robinson, G.O.; Vassar, R.; Sullivan, P.M.; Holtzman, D.M. Murine versus human apolipoprotein E4: Differential facilitation of and co-localization in cerebral amyloid angiopathy and amyloid plaques in APP transgenic mouse models. Acta Neuropathol. Commun. 2015, 3, 70. [Google Scholar] [CrossRef] [Green Version]
  138. Allcock, R.J.; Barrow, A.D.; Forbes, S.; Beck, S.; Trowsdale, J. The human TREM gene cluster at 6p21.1 encodes both activating and inhibitory single IgV domain receptors and includes NKp44. Eur. J. Immunol. 2003, 33, 567–577. [Google Scholar] [CrossRef]
  139. Chertoff, M.; Shrivastava, K.; Gonzalez, B.; Acarin, L.; Giménez-Llort, L. Differential modulation of TREM2 protein during postnatal brain development in mice. PLoS ONE 2013, 8, e72083. [Google Scholar] [CrossRef] [Green Version]
  140. Bhattacharjee, S.; Zhao, Y.; Dua, P.; Rogaev, E.I.; Lukiw, W.J. microRNA-34a-Mediated Down-Regulation of the Microglial-Enriched Triggering Receptor and Phagocytosis-Sensor TREM2 in Age-Related Macular Degeneration. PLoS ONE 2016, 11, e0150211. [Google Scholar] [CrossRef] [PubMed]
  141. Zhou, S.L.; Tan, C.C.; Hou, X.H.; Cao, X.P.; Tan, L.; Yu, J.T. TREM2 Variants and Neurodegenerative Diseases: A Systematic Review and Meta-Analysis. J. Alzheimers Dis. 2019, 68, 1171–1184. [Google Scholar] [CrossRef]
  142. Kober, D.L.; Alexander-Brett, J.M.; Karch, C.M.; Cruchaga, C.; Colonna, M.; Holtzman, M.J.; Brett, T.J. Neurodegenerative disease mutations in TREM2 reveal a functional surface and distinct loss-of-function mechanisms. Elife 2016, 5. [Google Scholar] [CrossRef]
  143. Kawabori, M.; Kacimi, R.; Kauppinen, T.; Calosing, C.; Kim, J.Y.; Hsieh, C.L.; Nakamura, M.C.; Yenari, M.A. Triggering receptor expressed on myeloid cells 2 (TREM2) deficiency attenuates phagocytic activities of microglia and exacerbates ischemic damage in experimental stroke. J. Neurosci. 2015, 35, 3384–3396. [Google Scholar] [CrossRef] [PubMed]
  144. Lue, L.F.; Schmitz, C.T.; Serrano, G.; Sue, L.I.; Beach, T.G.; Walker, D.G. TREM2 Protein Expression Changes Correlate with Alzheimer’s Disease Neurodegenerative Pathologies in Post-Mortem Temporal Cortices. Brain Pathol. 2015, 25, 469–480. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Perez, S.E.; Nadeem, M.; He, B.; Miguel, J.C.; Malek-Ahmadi, M.H.; Chen, K.; Mufson, E.J. Neocortical and hippocampal TREM2 protein levels during the progression of Alzheimer’s disease. Neurobiol. Aging 2017, 54, 133–143. [Google Scholar] [CrossRef] [PubMed]
  146. Jay, T.R.; Miller, C.M.; Cheng, P.J.; Graham, L.C.; Bemiller, S.; Broihier, M.L.; Xu, G.; Margevicius, D.; Karlo, J.C.; Sousa, G.L.; et al. TREM2 deficiency eliminates TREM2+ inflammatory macrophages and ameliorates pathology in Alzheimer’s disease mouse models. J. Exp. Med. 2015, 212, 287–295. [Google Scholar] [CrossRef] [PubMed]
  147. Wang, Y.; Cella, M.; Mallinson, K.; Ulrich, J.D.; Young, K.L.; Robinette, M.L.; Gilfillan, S.; Krishnan, G.M.; Sudhakar, S.; Zinselmeyer, B.H.; et al. TREM2 lipid sensing sustains the microglial response in an Alzheimer’s disease model. Cell 2015, 160, 1061–1071. [Google Scholar] [CrossRef] [Green Version]
  148. Wang, Y.; Ulland, T.K.; Ulrich, J.D.; Song, W.; Tzaferis, J.A.; Hole, J.T.; Yuan, P.; Mahan, T.E.; Shi, Y.; Gilfillan, S.; et al. TREM2-mediated early microglial response limits diffusion and toxicity of amyloid plaques. J. Exp. Med. 2016, 213, 667–675. [Google Scholar] [CrossRef] [Green Version]
  149. Jay, T.R.; Hirsch, A.M.; Broihier, M.L.; Miller, C.M.; Neilson, L.E.; Ransohoff, R.M.; Lamb, B.T.; Landreth, G.E. Disease Progression-Dependent Effects of TREM2 Deficiency in a Mouse Model of Alzheimer’s Disease. J. Neurosci. 2017, 37, 637–647. [Google Scholar] [CrossRef]
  150. Saito, T.; Matsuba, Y.; Mihira, N.; Takano, J.; Nilsson, P.; Itohara, S.; Iwata, N.; Saido, T.C. Single App knock-in mouse models of Alzheimer’s disease. Nat. Neurosci. 2014, 17, 661–663. [Google Scholar] [CrossRef]
  151. Pardossi-Piquard, R.; Checler, F. The physiology of the β-amyloid precursor protein intracellular domain AICD. J. Neurochem. 2012, 120 Suppl 1, 109–124. [Google Scholar] [CrossRef]
  152. Mitani, Y.; Yarimizu, J.; Saita, K.; Uchino, H.; Akashiba, H.; Shitaka, Y.; Ni, K.; Matsuoka, N. Differential effects between γ-secretase inhibitors and modulators on cognitive function in amyloid precursor protein-transgenic and nontransgenic mice. J. Neurosci. 2012, 32, 2037–2050. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Sasaguri, H.; Nilsson, P.; Hashimoto, S.; Nagata, K.; Saito, T.; De Strooper, B.; Hardy, J.; Vassar, R.; Winblad, B.; Saido, T.C. APP mouse models for Alzheimer’s disease preclinical studies. EMBO J. 2017, 36, 2473–2487. [Google Scholar] [CrossRef] [PubMed]
  154. Masuda, A.; Kobayashi, Y.; Kogo, N.; Saito, T.; Saido, T.C.; Itohara, S. Cognitive deficits in single App knock-in mouse models. Neurobiol. Learn. Mem. 2016, 135, 73–82. [Google Scholar] [CrossRef] [Green Version]
  155. Sakakibara, Y.; Sekiya, M.; Saito, T.; Saido, T.C.; Iijima, K.M. Cognitive and emotional alterations in App knock-in mouse models of Aβ amyloidosis. BMC Neurosci. 2018, 19, 46. [Google Scholar] [CrossRef] [PubMed]
  156. Whyte, L.S.; Hemsley, K.M.; Lau, A.A.; Hassiotis, S.; Saito, T.; Saido, T.C.; Hopwood, J.J.; Sargeant, T.J. Reduction in open field activity in the absence of memory deficits in the App(NL-G-F) knock-in mouse model of Alzheimer’s disease. Behav. Brain Res. 2018, 336, 177–181. [Google Scholar] [CrossRef] [PubMed]
  157. Latif-Hernandez, A.; Shah, D.; Craessaerts, K.; Saido, T.; Saito, T.; De Strooper, B.; Van der Linden, A.; D’Hooge, R. Subtle behavioral changes and increased prefrontal-hippocampal network synchronicity in APP(NL-G-F) mice before prominent plaque deposition. Behav. Brain Res. 2019, 364, 431–441. [Google Scholar] [CrossRef] [Green Version]
  158. De Roeck, E.E.; Engelborghs, S.; Dierckx, E. Next Generation Brain Health Depends on Early Alzheimer Disease Diagnosis: From a Timely Diagnosis to Future Population Screening. J. Am. Med. Dir. Assoc. 2016, 17, 452–453. [Google Scholar] [CrossRef] [PubMed]
  159. Farlow, M.; Anand, R.; Messina, J., Jr.; Hartman, R.; Veach, J. A 52-week study of the efficacy of rivastigmine in patients with mild to moderately severe Alzheimer’s disease. Eur. Neurol. 2000, 44, 236–241. [Google Scholar] [CrossRef] [PubMed]
  160. Doraiswamy, P.M.; Krishnan, K.R.; Anand, R.; Sohn, H.; Danyluk, J.; Hartman, R.D.; Veach, J. Long-term effects of rivastigmine in moderately severe Alzheimer’s disease: Does early initiation of therapy offer sustained benefits? Prog. Neuropsychopharmacol. Biol. Psychiatry 2002, 26, 705–712. [Google Scholar] [CrossRef]
  161. Romberg, C.; Mattson, M.P.; Mughal, M.R.; Bussey, T.J.; Saksida, L.M. Impaired attention in the 3xTgAD mouse model of Alzheimer’s disease: Rescue by donepezil (Aricept). J. Neurosci. 2011, 31, 3500–3507. [Google Scholar] [CrossRef] [Green Version]
  162. Bussey, T.J.; Holmes, A.; Lyon, L.; Mar, A.C.; McAllister, K.A.; Nithianantharajah, J.; Oomen, C.A.; Saksida, L.M. New translational assays for preclinical modelling of cognition in schizophrenia: The touchscreen testing method for mice and rats. Neuropharmacology 2012, 62, 1191–1203. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Romberg, C.; Bussey, T.J.; Saksida, L.M. Paying more attention to attention: Towards more comprehensive cognitive translation using mouse models of Alzheimer’s disease. Brain Res. Bull. 2013, 92, 49–55. [Google Scholar] [CrossRef] [PubMed]
  164. Saifullah, M.A.B.; Komine, O.; Dong, Y.; Fukumoto, K.; Sobue, A.; Endo, F.; Saito, T.; Saido, T.C.; Yamanaka, K.; Mizoguchi, H. Touchscreen-based location discrimination and paired associate learning tasks detect cognitive impairment at an early stage in an App knock-in mouse model of Alzheimer’s disease. Mol. Brain 2020, 13, 147. [Google Scholar] [CrossRef] [PubMed]
  165. Sobue, A.; Komine, O.; Hara, Y.; Endo, F.; Mizoguchi, H.; Watanabe, S.; Murayama, S.; Saito, T.; Saido, T.C.; Sahara, N.; et al. Microglial gene signature reveals loss of homeostatic microglia associated with neurodegeneration of Alzheimer’s disease. Acta Neuropathol. Commun. 2021, 9, 1. [Google Scholar] [CrossRef]
  166. Kim, K.Y.; Suh, Y.H.; Chang, K.A. Therapeutic Effects of Human Amniotic Epithelial Stem Cells in a Transgenic Mouse Model of Alzheimer’s Disease. Int. J. Mol. Sci. 2020, 21, 2658. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Schmid, S.; Rammes, G.; Blobner, M.; Kellermann, K.; Bratke, S.; Fendl, D.; Kaichuan, Z.; Schneider, G.; Jungwirth, B. Cognitive decline in Tg2576 mice shows sex-specific differences and correlates with cerebral amyloid-beta. Behav. Brain Res. 2019, 359, 408–417. [Google Scholar] [CrossRef]
  168. Evans, C.; Hvoslef-Eide, M.; Thomas, R.; Kidd, E.; Good, M.A. A rapidly acquired foraging-based working memory task, sensitive to hippocampal lesions, reveals age-dependent and age-independent behavioural changes in a mouse model of amyloid pathology. Neurobiol. Learn. Mem. 2018, 149, 46–57. [Google Scholar] [CrossRef] [PubMed]
  169. Lim, S.L.; Tran, D.N.; Kieu, Z.; Chen, C.; Villanueva, E.; Ghiaar, S.; Gallup, V.; Zumkehr, J.; Cribbs, D.H.; Rodriguez-Ortiz, C.J.; et al. Genetic Ablation of Hematopoietic Cell Kinase Accelerates Alzheimer’s Disease-Like Neuropathology in Tg2576 Mice. Mol. Neurobiol. 2020, 57, 2447–2460. [Google Scholar] [CrossRef]
  170. Tanaka, T.; Hirai, S.; Hosokawa, M.; Saito, T.; Sakuma, H.; Saido, T.; Hasegawa, M.; Okado, H. Early-life stress induces the development of Alzheimer’s disease pathology via angiopathy. Exp. Neurol. 2021, 337, 113552. [Google Scholar] [CrossRef]
  171. Chun, Y.S.; Zhang, L.; Li, H.; Park, Y.; Chung, S.; Yang, H.O. 7-Deoxy-trans-dihydronarciclasine Reduces β-Amyloid and Ameliorates Memory Impairment in a Transgenic Model of Alzheimer’s Disease. Mol. Neurobiol. 2018, 55, 8953–8964. [Google Scholar] [CrossRef]
  172. Pignataro, A.; Meli, G.; Pagano, R.; Fontebasso, V.; Battistella, R.; Conforto, G.; Ammassari-Teule, M.; Middei, S. Activity-Induced Amyloid-β Oligomers Drive Compensatory Synaptic Rearrangements in Brain Circuits Controlling Memory of Presymptomatic Alzheimer’s Disease Mice. Biol. Psychiatry 2019, 86, 185–195. [Google Scholar] [CrossRef] [PubMed]
  173. Vilella, A.; Belletti, D.; Sauer, A.K.; Hagmeyer, S.; Sarowar, T.; Masoni, M.; Stasiak, N.; Mulvihill, J.J.E.; Ruozi, B.; Forni, F.; et al. Reduced plaque size and inflammation in the APP23 mouse model for Alzheimer’s disease after chronic application of polymeric nanoparticles for CNS targeted zinc delivery. J. Trace Elem. Med. Biol. 2018, 49, 210–221. [Google Scholar] [CrossRef] [PubMed]
  174. Wang, X.; Liu, D.; Huang, H.Z.; Wang, Z.H.; Hou, T.Y.; Yang, X.; Pang, P.; Wei, N.; Zhou, Y.F.; Dupras, M.J.; et al. A Novel MicroRNA-124/PTPN1 Signal Pathway Mediates Synaptic and Memory Deficits in Alzheimer’s Disease. Biol. Psychiatry 2018, 83, 395–405. [Google Scholar] [CrossRef] [PubMed]
  175. Elhaik Goldman, S.; Goez, D.; Last, D.; Naor, S.; Liraz Zaltsman, S.; Sharvit-Ginon, I.; Atrakchi-Baranes, D.; Shemesh, C.; Twitto-Greenberg, R.; Tsach, S.; et al. High-fat diet protects the blood-brain barrier in an Alzheimer’s disease mouse model. Aging Cell 2018, 17, e12818. [Google Scholar] [CrossRef]
  176. Dong, H.; Csernansky, C.A.; Martin, M.V.; Bertchume, A.; Vallera, D.; Csernansky, J.G. Acetylcholinesterase inhibitors ameliorate behavioral deficits in the Tg2576 mouse model of Alzheimer’s disease. Psychopharmacology (Berl.) 2005, 181, 145–152. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Unger, C.; Svedberg, M.M.; Yu, W.F.; Hedberg, M.M.; Nordberg, A. Effect of subchronic treatment of memantine, galantamine, and nicotine in the brain of Tg2576 (APPswe) transgenic mice. J. Pharmacol. Exp. Ther. 2006, 317, 30–36. [Google Scholar] [CrossRef]
  178. Dong, H.; Yuede, C.M.; Coughlan, C.; Lewis, B.; Csernansky, J.G. Effects of memantine on neuronal structure and conditioned fear in the Tg2576 mouse model of Alzheimer’s disease. Neuropsychopharmacology 2008, 33, 3226–3236. [Google Scholar] [CrossRef]
  179. Dong, H.; Yuede, C.M.; Coughlan, C.A.; Murphy, K.M.; Csernansky, J.G. Effects of donepezil on amyloid-beta and synapse density in the Tg2576 mouse model of Alzheimer’s disease. Brain Res. 2009, 1303, 169–178. [Google Scholar] [CrossRef] [Green Version]
  180. Van Dam, D.; Abramowski, D.; Staufenbiel, M.; De Deyn, P.P. Symptomatic effect of donepezil, rivastigmine, galantamine and memantine on cognitive deficits in the APP23 model. Psychopharmacology (Berlin) 2005, 180, 177–190. [Google Scholar] [CrossRef]
  181. Rogers, S.L.; Doody, R.S.; Mohs, R.C.; Friedhoff, L.T. Donepezil improves cognition and global function in Alzheimer disease: A 15-week, double-blind, placebo-controlled study. Donepezil Study Group. Arch. Intern. Med. 1998, 158, 1021–1031. [Google Scholar] [CrossRef]
  182. Winblad, B.; Cummings, J.; Andreasen, N.; Grossberg, G.; Onofrj, M.; Sadowsky, C.; Zechner, S.; Nagel, J.; Lane, R. A six-month double-blind, randomized, placebo-controlled study of a transdermal patch in Alzheimer’s disease--rivastigmine patch versus capsule. Int. J. Geriatr. Psychiatry 2007, 22, 456–467. [Google Scholar] [CrossRef] [PubMed]
  183. Kavanagh, S.; Van Baelen, B.; Schäuble, B. Long-term effects of galantamine on cognitive function in Alzheimer’s disease: A large-scale international retrospective study. J. Alzheimers Dis. 2011, 27, 521–530. [Google Scholar] [CrossRef] [PubMed]
  184. Wilkinson, D.; Fox, N.C.; Barkhof, F.; Phul, R.; Lemming, O.; Scheltens, P. Memantine and brain atrophy in Alzheimer’s disease: A 1-year randomized controlled trial. J. Alzheimers Dis. 2012, 29, 459–469. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Lim, G.P.; Yang, F.; Chu, T.; Chen, P.; Beech, W.; Teter, B.; Tran, T.; Ubeda, O.; Ashe, K.H.; Frautschy, S.A.; et al. Ibuprofen suppresses plaque pathology and inflammation in a mouse model for Alzheimer’s disease. J. Neurosci. 2000, 20, 5709–5714. [Google Scholar] [CrossRef]
  186. Kotilinek, L.A.; Westerman, M.A.; Wang, Q.; Panizzon, K.; Lim, G.P.; Simonyi, A.; Lesne, S.; Falinska, A.; Younkin, L.H.; Younkin, S.G.; et al. Cyclooxygenase-2 inhibition improves amyloid-beta-mediated suppression of memory and synaptic plasticity. Brain 2008, 131, 651–664. [Google Scholar] [CrossRef]
  187. McKee, A.C.; Carreras, I.; Hossain, L.; Ryu, H.; Klein, W.L.; Oddo, S.; LaFerla, F.M.; Jenkins, B.G.; Kowall, N.W.; Dedeoglu, A. Ibuprofen reduces Abeta, hyperphosphorylated tau and memory deficits in Alzheimer mice. Brain Res. 2008, 1207, 225–236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Hillmann, A.; Hahn, S.; Schilling, S.; Hoffmann, T.; Demuth, H.U.; Bulic, B.; Schneider-Axmann, T.; Bayer, T.A.; Weggen, S.; Wirths, O. No improvement after chronic ibuprofen treatment in the 5XFAD mouse model of Alzheimer’s disease. Neurobiol. Aging 2012, 33, 833.e839–850. [Google Scholar] [CrossRef]
  189. Babiloni, C.; Frisoni, G.B.; Del Percio, C.; Zanetti, O.; Bonomini, C.; Cassetta, E.; Pasqualetti, P.; Miniussi, C.; De Rosas, M.; Valenzano, A.; et al. Ibuprofen treatment modifies cortical sources of EEG rhythms in mild Alzheimer’s disease. Clin. Neurophysiol. 2009, 120, 709–718. [Google Scholar] [CrossRef] [Green Version]
  190. Pasqualetti, P.; Bonomini, C.; Dal Forno, G.; Paulon, L.; Sinforiani, E.; Marra, C.; Zanetti, O.; Rossini, P.M. A randomized controlled study on effects of ibuprofen on cognitive progression of Alzheimer’s disease. Aging Clin. Exp. Res. 2009, 21, 102–110. [Google Scholar] [CrossRef]
  191. Nenov, M.N.; Laezza, F.; Haidacher, S.J.; Zhao, Y.; Sadygov, R.G.; Starkey, J.M.; Spratt, H.; Luxon, B.A.; Dineley, K.T.; Denner, L. Cognitive enhancing treatment with a PPARγ agonist normalizes dentate granule cell presynaptic function in Tg2576 APP mice. J. Neurosci. 2014, 34, 1028–1036. [Google Scholar] [CrossRef] [Green Version]
  192. O’Reilly, J.A.; Lynch, M. Rosiglitazone improves spatial memory and decreases insoluble Aβ(1-42) in APP/PS1 mice. J. Neuroimmune Pharmacol. 2012, 7, 140–144. [Google Scholar] [CrossRef]
  193. Searcy, J.L.; Phelps, J.T.; Pancani, T.; Kadish, I.; Popovic, J.; Anderson, K.L.; Beckett, T.L.; Murphy, M.P.; Chen, K.C.; Blalock, E.M.; et al. Long-term pioglitazone treatment improves learning and attenuates pathological markers in a mouse model of Alzheimer’s disease. J. Alzheimers Dis. 2012, 30, 943–961. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Gold, M.; Alderton, C.; Zvartau-Hind, M.; Egginton, S.; Saunders, A.M.; Irizarry, M.; Craft, S.; Landreth, G.; Linnamägi, U.; Sawchak, S. Rosiglitazone monotherapy in mild-to-moderate Alzheimer’s disease: Results from a randomized, double-blind, placebo-controlled phase III study. Dement. Geriatr. Cogn. Disord. 2010, 30, 131–146. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Tzimopoulou, S.; Cunningham, V.J.; Nichols, T.E.; Searle, G.; Bird, N.P.; Mistry, P.; Dixon, I.J.; Hallett, W.A.; Whitcher, B.; Brown, A.P.; et al. A multi-center randomized proof-of-concept clinical trial applying [¹⁸F]FDG-PET for evaluation of metabolic therapy with rosiglitazone XR in mild to moderate Alzheimer’s disease. J. Alzheimers Dis. 2010, 22, 1241–1256. [Google Scholar] [CrossRef] [PubMed]
  196. Harrington, C.; Sawchak, S.; Chiang, C.; Davies, J.; Donovan, C.; Saunders, A.M.; Irizarry, M.; Jeter, B.; Zvartau-Hind, M.; van Dyck, C.H.; et al. Rosiglitazone does not improve cognition or global function when used as adjunctive therapy to AChE inhibitors in mild-to-moderate Alzheimer’s disease: Two phase 3 studies. Curr. Alzheimer Res. 2011, 8, 592–606. [Google Scholar] [CrossRef] [PubMed]
  197. Miller, B.W.; Willett, K.C.; Desilets, A.R. Rosiglitazone and pioglitazone for the treatment of Alzheimer’s disease. Ann. Pharmacother. 2011, 45, 1416–1424. [Google Scholar] [CrossRef] [PubMed]
  198. Li, G.; Mayer, C.L.; Morelli, D.; Millard, S.P.; Raskind, W.H.; Petrie, E.C.; Cherrier, M.; Fagan, A.M.; Raskind, M.A.; Peskind, E.R. Effect of simvastatin on CSF Alzheimer disease biomarkers in cognitively normal adults. Neurology 2017, 89, 1251–1255. [Google Scholar] [CrossRef]
  199. Umeda, T.; Ono, K.; Sakai, A.; Yamashita, M.; Mizuguchi, M.; Klein, W.L.; Yamada, M.; Mori, H.; Tomiyama, T. Rifampicin is a candidate preventive medicine against amyloid-β and tau oligomers. Brain 2016, 139, 1568–1586. [Google Scholar] [CrossRef] [Green Version]
  200. Fukumoto, H.; Takahashi, H.; Tarui, N.; Matsui, J.; Tomita, T.; Hirode, M.; Sagayama, M.; Maeda, R.; Kawamoto, M.; Hirai, K.; et al. A noncompetitive BACE1 inhibitor TAK-070 ameliorates Abeta pathology and behavioral deficits in a mouse model of Alzheimer’s disease. J. Neurosci. 2010, 30, 11157–11166. [Google Scholar] [CrossRef]
  201. May, P.C.; Dean, R.A.; Lowe, S.L.; Martenyi, F.; Sheehan, S.M.; Boggs, L.N.; Monk, S.A.; Mathes, B.M.; Mergott, D.J.; Watson, B.M.; et al. Robust central reduction of amyloid-β in humans with an orally available, non-peptidic β-secretase inhibitor. J. Neurosci. 2011, 31, 16507–16516. [Google Scholar] [CrossRef] [Green Version]
  202. Elvang, A.B.; Volbracht, C.; Pedersen, L.; Jensen, K.G.; Karlsson, J.J.; Larsen, S.A.; Mørk, A.; Stensbøl, T.B.; Bastlund, J.F. Differential effects of gamma-secretase and BACE1 inhibition on brain Abeta levels in vitro and in vivo. J. Neurochem. 2009, 110, 1377–1387. [Google Scholar] [CrossRef] [PubMed]
  203. Extance, A. Alzheimer’s failure raises questions about disease-modifying strategies. Nat. Rev. Drug Discov. 2010, 9, 749–751. [Google Scholar] [CrossRef] [PubMed]
  204. Doody, R.S.; Raman, R.; Farlow, M.; Iwatsubo, T.; Vellas, B.; Joffe, S.; Kieburtz, K.; He, F.; Sun, X.; Thomas, R.G.; et al. A phase 3 trial of semagacestat for treatment of Alzheimer’s disease. N. Engl. J. Med. 2013, 369, 341–350. [Google Scholar] [CrossRef] [PubMed]
  205. Egan, M.F.; Kost, J.; Tariot, P.N.; Aisen, P.S.; Cummings, J.L.; Vellas, B.; Sur, C.; Mukai, Y.; Voss, T.; Furtek, C.; et al. Randomized Trial of Verubecestat for Mild-to-Moderate Alzheimer’s Disease. N. Engl. J. Med. 2018, 378, 1691–1703. [Google Scholar] [CrossRef] [PubMed]
  206. Schenk, D.; Barbour, R.; Dunn, W.; Gordon, G.; Grajeda, H.; Guido, T.; Hu, K.; Huang, J.; Johnson-Wood, K.; Khan, K.; et al. Immunization with amyloid-beta attenuates Alzheimer-disease-like pathology in the PDAPP mouse. Nature 1999, 400, 173–177. [Google Scholar] [CrossRef] [PubMed]
  207. Morgan, D.; Diamond, D.M.; Gottschall, P.E.; Ugen, K.E.; Dickey, C.; Hardy, J.; Duff, K.; Jantzen, P.; DiCarlo, G.; Wilcock, D.; et al. A beta peptide vaccination prevents memory loss in an animal model of Alzheimer’s disease. Nature 2000, 408, 982–985. [Google Scholar] [CrossRef]
  208. Kotilinek, L.A.; Bacskai, B.; Westerman, M.; Kawarabayashi, T.; Younkin, L.; Hyman, B.T.; Younkin, S.; Ashe, K.H. Reversible memory loss in a mouse transgenic model of Alzheimer’s disease. J. Neurosci. 2002, 22, 6331–6335. [Google Scholar] [CrossRef]
  209. Wilcock, D.M.; DiCarlo, G.; Henderson, D.; Jackson, J.; Clarke, K.; Ugen, K.E.; Gordon, M.N.; Morgan, D. Intracranially administered anti-Abeta antibodies reduce beta-amyloid deposition by mechanisms both independent of and associated with microglial activation. J. Neurosci. 2003, 23, 3745–3751. [Google Scholar] [CrossRef]
  210. Bard, F.; Cannon, C.; Barbour, R.; Burke, R.L.; Games, D.; Grajeda, H.; Guido, T.; Hu, K.; Huang, J.; Johnson-Wood, K.; et al. Peripherally administered antibodies against amyloid beta-peptide enter the central nervous system and reduce pathology in a mouse model of Alzheimer disease. Nat. Med. 2000, 6, 916–919. [Google Scholar] [CrossRef] [Green Version]
  211. Dodart, J.C.; Bales, K.R.; Gannon, K.S.; Greene, S.J.; DeMattos, R.B.; Mathis, C.; DeLong, C.A.; Wu, S.; Wu, X.; Holtzman, D.M.; et al. Immunization reverses memory deficits without reducing brain Abeta burden in Alzheimer’s disease model. Nat. Neurosci. 2002, 5, 452–457. [Google Scholar] [CrossRef] [PubMed]
  212. Gilman, S.; Koller, M.; Black, R.S.; Jenkins, L.; Griffith, S.G.; Fox, N.C.; Eisner, L.; Kirby, L.; Rovira, M.B.; Forette, F.; et al. Clinical effects of Abeta immunization (AN1792) in patients with AD in an interrupted trial. Neurology 2005, 64, 1553–1562. [Google Scholar] [CrossRef]
  213. Vellas, B.; Black, R.; Thal, L.J.; Fox, N.C.; Daniels, M.; McLennan, G.; Tompkins, C.; Leibman, C.; Pomfret, M.; Grundman, M. Long-term follow-up of patients immunized with AN1792: Reduced functional decline in antibody responders. Curr. Alzheimer Res. 2009, 6, 144–151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Delnomdedieu, M.; Duvvuri, S.; Li, D.J.; Atassi, N.; Lu, M.; Brashear, H.R.; Liu, E.; Ness, S.; Kupiec, J.W. First-In-Human safety and long-term exposure data for AAB-003 (PF-05236812) and biomarkers after intravenous infusions of escalating doses in patients with mild to moderate Alzheimer’s disease. Alzheimers Res. Ther. 2016, 8, 12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Ostrowitzki, S.; Lasser, R.A.; Dorflinger, E.; Scheltens, P.; Barkhof, F.; Nikolcheva, T.; Ashford, E.; Retout, S.; Hofmann, C.; Delmar, P.; et al. A phase III randomized trial of gantenerumab in prodromal Alzheimer’s disease. Alzheimers Res. Ther. 2017, 9, 95. [Google Scholar] [CrossRef]
  216. Honig, L.S.; Vellas, B.; Woodward, M.; Boada, M.; Bullock, R.; Borrie, M.; Hager, K.; Andreasen, N.; Scarpini, E.; Liu-Seifert, H.; et al. Trial of Solanezumab for Mild Dementia Due to Alzheimer’s Disease. N. Engl. J. Med. 2018, 378, 321–330. [Google Scholar] [CrossRef] [PubMed]
  217. Tolar, M.; Abushakra, S.; Hey, J.A.; Porsteinsson, A.; Sabbagh, M. Aducanumab, gantenerumab, BAN2401, and ALZ-801-the first wave of amyloid-targeting drugs for Alzheimer’s disease with potential for near term approval. Alzheimers Res. Ther. 2020, 12, 95. [Google Scholar] [CrossRef]
  218. Vandenberghe, R.; Riviere, M.E.; Caputo, A.; Sovago, J.; Maguire, R.P.; Farlow, M.; Marotta, G.; Sanchez-Valle, R.; Scheltens, P.; Ryan, J.M.; et al. Active Aβ immunotherapy CAD106 in Alzheimer’s disease: A phase 2b study. Alzheimers Dement (N Y). 2017, 3, 10–22. [Google Scholar] [CrossRef]
  219. van Dyck, C.H. Anti-Amyloid-β Monoclonal Antibodies for Alzheimer’s Disease: Pitfalls and Promise. Biol. Psychiatry 2018, 83, 311–319. [Google Scholar] [CrossRef] [Green Version]
  220. Sevigny, J.; Chiao, P.; Bussière, T.; Weinreb, P.H.; Williams, L.; Maier, M.; Dunstan, R.; Salloway, S.; Chen, T.; Ling, Y.; et al. The antibody aducanumab reduces Aβ plaques in Alzheimer’s disease. Nature 2016, 537, 50–56. [Google Scholar] [CrossRef]
  221. Theunis, C.; Crespo-Biel, N.; Gafner, V.; Pihlgren, M.; López-Deber, M.P.; Reis, P.; Hickman, D.T.; Adolfsson, O.; Chuard, N.; Ndao, D.M.; et al. Efficacy and safety of a liposome-based vaccine against protein Tau, assessed in tau.P301L mice that model tauopathy. PLoS ONE 2013, 8, e72301. [Google Scholar] [CrossRef] [Green Version]
  222. Novak, P.; Schmidt, R.; Kontsekova, E.; Kovacech, B.; Smolek, T.; Katina, S.; Fialova, L.; Prcina, M.; Parrak, V.; Dal-Bianco, P.; et al. FUNDAMANT: An interventional 72-week phase 1 follow-up study of AADvac1, an active immunotherapy against tau protein pathology in Alzheimer’s disease. Alzheimers Res. Ther. 2018, 10, 108. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Table 1. Summary of representative biomarkers in AD.
Table 1. Summary of representative biomarkers in AD.
ModalityTypeReference
CSF[36]
Tau[38,39,40]
Aβ oligomers[42,43]
PETAmyloid[44,45,46,47]
Tau[48,49]
Microglia[50,51]
Blood[55,56,57,58]
Secondary structure of Aβ[59]
Tau[60,61,62]
IL-8[63]
Neurofilament light[64]
MicroRNAs[65]
IGF-1 and IGF-binding protein-3[66]
APOA-I, C3, and transthyretin[67]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nakai, T.; Yamada, K.; Mizoguchi, H. Alzheimer’s Disease Animal Models: Elucidation of Biomarkers and Therapeutic Approaches for Cognitive Impairment. Int. J. Mol. Sci. 2021, 22, 5549. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22115549

AMA Style

Nakai T, Yamada K, Mizoguchi H. Alzheimer’s Disease Animal Models: Elucidation of Biomarkers and Therapeutic Approaches for Cognitive Impairment. International Journal of Molecular Sciences. 2021; 22(11):5549. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22115549

Chicago/Turabian Style

Nakai, Tsuyoshi, Kiyofumi Yamada, and Hiroyuki Mizoguchi. 2021. "Alzheimer’s Disease Animal Models: Elucidation of Biomarkers and Therapeutic Approaches for Cognitive Impairment" International Journal of Molecular Sciences 22, no. 11: 5549. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22115549

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop