Next Article in Journal
GPR68: An Emerging Drug Target in Cancer
Next Article in Special Issue
Recent Developments in Metal-Based Drugs and Chelating Agents for Neurodegenerative Diseases Treatments
Previous Article in Journal
Mechanisms Underlying the Visual Benefit of Cell Transplantation for the Treatment of Retinal Degenerations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Drug Development for Alzheimer’s Disease: Microglia Induced Neuroinflammation as a Target?

1
Department of Biochemistry, Medical College, Qingdao University, Qingdao 266071, China
2
Beijing Institute for Brain Disorders, Capital Medical University, Beijing 100069, China
3
The Brain Science Center, Beijing Institute of Basic Medical Sciences, 100850 Beijing, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2019, 20(3), 558; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms20030558
Submission received: 27 December 2018 / Revised: 9 January 2019 / Accepted: 13 January 2019 / Published: 28 January 2019

Abstract

:
Alzheimer’s disease (AD) is one of the most common causes of dementia. Its pathogenesis is characterized by the aggregation of the amyloid-β (Aβ) protein in senile plaques and the hyperphosphorylated tau protein in neurofibrillary tangles in the brain. Current medications for AD can provide temporary help with the memory symptoms and other cognitive changes of patients, however, they are not able to stop or reverse the progression of AD. New medication discovery and the development of a cure for AD is urgently in need. In this review, we summarized drugs for AD treatments and their recent updates, and discussed the potential of microglia induced neuroinflammation as a target for anti-AD drug development.

1. Introduction

Alzheimer’s disease (AD) is a chronic neurodegenerative disorder, characterized by a gradually progressive loss of memory and cognitive functions as early symptoms, and developing into dementia eventually [1]. It is mostly diagnosed in people over 65 years-old, which is termed sporadic AD, while around 4–5% of cases occur before 65, which is classified as early-onset AD [2]. According to the recent report released by Alzheimer’s Disease International (ADI), AD has become one of the most common causes of dementia. In 2018, 50 million people are suffering from dementia, costing 1 trillion US$ globally. By 2050, the estimated number of people with dementia will reach 152 million, causing a huge social and economic burden for the families and caregivers of the patients. Incidence of AD is sex-related, which happens in women more than men [3,4]. In the United States, among the 5.5 million patients diagnosed with sporadic AD, 3.4 million are women, which makes women almost twice more vulnerable than men [5]. Multiple causes may explain this higher incidence of AD in women, including the difference of life expectancy [6], sex steroid hormones [7,8,9], and educational level [10,11] of men and women.
It has been more than a century since the first diagnosis of Alzheimer’s disease in 1906 [12], and the cause of this disease is still unclear. Consequently, pharmacological approaches to treat AD are mostly symptomatic. Currently, no drug is able to stop or reverse the progression of AD. In recent decades, amyloid-β (Aβ) plaques and tau neurofibrillary tangles aggregations have been intensively studied, and are believed to be vital targets for the cure of AD. Many new drugs have been developed and have entered clinical trials. However, up until now, no Aβ-targeting drug has been officially approved by the United States Food and Drug Administration (FDA) for the clinical treatment of AD.
Microglia-mediated neuroinflammation is one of the most remarkable hallmarks in neurodegenerative diseases. Microglia induced neuroinflammation contributes to the pathogenesis of AD by direct damage to the neuron, concurrently promoting protein aggregations, suggesting that it should be a new target for AD treatment [13]. In this review, we summarized the Aβ plaques and tau neurofibrillary tangles-targeting drugs currently undergoing clinical trials (information comes from https://clinicaltrials.gov), and discussed the potential of microglia induced neuroinflammation as a target for anti-AD drug development.

2. Cause of Alzheimer’s Disease

The pathology of AD includes the aggregation of extracellular senile plaques formed by Aβ protein, intracellular neurofibrillary tangles formed by hyperphosphorylated tau protein, enhanced neuroinflammation, oxidative stress, iron dysregulation, and neuronal cell death [14,15,16]. The symptoms of AD patients usually develop starting from mild cognitive impairment (MCI) at the preclinical stage, to the complete loss of language and the ability to live independently at the advanced stage. Multiple hypotheses exist trying to explain the pathogenesis of AD, including cholinergic hypothesis, amyloid cascade hypothesis, tau neurofibrillary hypothesis, mitochondrial dysfunction, and so on.
While AD is not considered a genetically inherited disease, mutations in the genes encoding the Amyloid precursor protein (APP), presenilins 1 and 2, can cause familial AD, usually with an early onset [17,18]. Apolipoprotein E (ApoE) ε4 allele is the best known genetic risk factor in the incidence of sporadic AD [1,16,19]. Individuals with ApoE ε4/ε4 genotypes have significantly increased incidences of AD compared with individuals with the ApoE ε3/ε4 genotypes [20]. Although no difference in the incidence of AD is observed between men and women of the ages between 55 to 58, women show a higher risk at an earlier age [20]. Mutations in the gene encoding the triggering receptor expressed on myeloid cells 2 (TREM2) are also proven to increase the risk of AD [21,22,23,24,25]. A TREM2 variant, rs75932628, results in an Arg47His substitution, significantly increasing the incidence of AD [21,22]. Calcium (Ca+), as a universal second messenger, involves in a wide range of cellular processes. Neural Ca+ dysfunction has been widely accepted as an important contributor in AD and other neurodegenerative diseases [26,27,28]. Functional intracellular calcium homeostasis is tightly regulated within a narrow range by Ca+ channels and pumps [29,30]. Calcium homeostasis modulator protein 1 (CALHM1) plays important roles in controlling the Ca+ influx and intracellular calcium signaling, through the activation of extracellular signal-regulated kinase-1/-2 (ERK1/2) kinase signaling cascade [31,32]. CALHM1 knocked out mice displayed an impaired memory flexibility and hippocampal long-term potentiation (LTP), indicating Ca+ dysregulation as an important factor in neuronal activity [32]. Other causes of AD include metal ions dysregulation and mitochondrial dysfunction related to protein aggregations, oxidative stress, and neuron death [33,34,35,36,37]. A recent meta-analysis of genome-wide association studies (GWAS) identified 19 other loci associated with AD as genetic risk factors [38].
The cholinergic hypothesis proposes that AD is caused by reduced neurotransmitter acetylcholine synthesis [39]. It is the first hypothesis established that tries to explain the onset and development of AD, and is the most important target that current clinical treatments are based on [39,40]. However, this hypothesis is still controversial, largely because of the questionable efficacy of the anti-AD drugs intended to treat acetylcholine deficiency [40].
In the amyloid cascade hypothesis, senile plaques aggregation at the extracellular region of the human brain is responsible for an amyloid neurotoxic cascade, resulting in the atrophy and degeneration in the temporal and parietal lobe, pre-frontal cortex, and hippocampus, thus causing memory and cognitive impairment, eventually developing into dementia [41,42]. Aβ protein, the core component of senile plaques, is the product of the sequential cleavage of transmembrane APP by β-secretase and γ-secretase [43,44]. In a non-amyloidogenic pathway, APP is processed by α-secretase instead of β-secretase 1 to form soluble amyloid precursor protein-α (sAPP-α), and it yields P3 as the final product [45]. Transgenic mice carrying the mutant human APP gene are able to develop AD-like symptoms, such as amyloid plaques and spatial learning deficits [46,47]. Meanwhile, ApoE enhances the break-down of the Aβ protein. The isoform ApoE ε4 is not effective in this reaction, leading to increased piled up of Aβ in the extracellular region [48]. This hypothesis is supported by the finding that 40–65% of AD patients carry at least one ApoE ε4 allele [49].
Aβ toxicity to the neuron cells and the ability to magnify its effect through positive feedback loop is the main focus of this hypothesis. Aβ accumulates at synapse, causing impairment in the synaptic functions as well as neurotransmission. Also, Aβ accumulation at the extracellular region is able to cause not only neuronal cell death, but also the loss of postsynaptic density protein 95 (PSD-95) and synaptophysin [50]. The excitotoxicity of Aβ is due to the over-excitation of N-methyl-d-aspartate receptors (NMDARs), and is considered as the main mechanism of Aβ-induced neuron damage [51]. Apart from that, Aβ is able to induce hyperphosphorylated tau protein accumulation, forming fibril tangles at the intracellular region [52], and causes the chronic neuroinflammation. Multiple forms of Aβ exist in the extracellular region of the AD brain. Aβ40 and Aβ42 are the most common isoforms of this protein, both of which are found in the amyloid plaques in the brains of AD patients. Aβ42 has a particularly strong tendency to aggregate, giving rise to neurotoxic components, including oligomers and plaques. In the familial AD, many genetic mutations in APP and presenilins contribute to the disease by only increasing one of the Aβ isoforms, or by just altering the Aβ40/42 ratio [53]. Furthermore, substantial evidence now indicates that soluble Aβ oligomer accumulation is corelated with the progress of AD, rather than insoluble Aβ plaques aggregation [54,55]. Apart from Aβ itself, by-products of Aβ generation may also be able to contribute to AD’s progression. The cleaved N-terminal fragment of APP (N-APP) is reported to be able to bind to death receptor 6 (DR6), and initiates the degeneration of the neuron cells [56]. The amyloid cascade hypothesis has long since been proposed, and has been intensively studied for decades [57,58]; however, it has recently been contested, mainly because of the limited outcomes of the drugs targeting it in clinical trials [59,60,61].
Another well studied hypothesis is the Tau hypothesis, which proposes that the hyperphosphorylated tau protein aggregates to form neurofibrillary tangles at the intracellular space, which contributes to the cascade of the disease [62]. Tau protein is the stabilizer of microtubules, and it plays vital roles in the cell transport system, thus its abnormality affects the functions and synaptic transmission of AD [63]. Of more importance, the hyperphosphorylation of tau protein causes the collapse of microtubules, leads to a cascade disintegration, and eventually destroys the fundamental structure of the neuron’s transport system, which end ups with neuronal cell death [64,65].

3. Microglia Induced Neuroinflammation in Alzheimer’s Disease

Microglial cells are located throughout the central nerve system (CNS), and account for around 15% of the total cellular population in the brain [66]. Functioning as the resident macrophages, they are critical components in the immune defense and homeostasis maintenance of the CNS [67,68,69]. Microglia scavenges throughout the brain, activated by the presence of pathogenic invasion, tissue injury, and protein aggregates, and through receptors, it recognizes danger-associated molecular patterns (DAMPs) or pathogen-associated molecular patterns (PAMPs) [68,69,70]. Upon activation, microglia is able to migrate to the site of injury and initiate an innate immune response. Simultaneously, microglia is an important component in the protection and remodeling of synapse plasticity, which is important for the normal functions of neuronal circuits [71]. Furthermore, microglia plays important roles in learning and memory functions, through the protection and remodeling of learning-related synapse [72]. Using single-cell RNA sequencing, a novel subtype of microglia, disease-associated microglia (DAM), associated with the AD progression in both mouse models and human patients, is discovered [73], indicating the vital roles of the microglia and its dynamic in the progression of AD. Generally, DAM displays a unique two-step activation mechanism along with the progression of AD; it starts with an initial TREM2-independent activation, which involves changes in the microglia markers and genes associated with AD, followed by a second TREM2-dependent activation characterized by expressing high levels of lipid metabolism and phagocytic genes signatures [73]. As the most important factor correlated with AD, aging, especially the aging of the microglia, is a vital factor in the progression of AD. Aged microglia expresses high levels of pro-inflammatory cytokines, including interleukin-1beta (IL-1β), tumor necrosis factor-alpha (TNF-α), and IL-6 [74], and undergoes a series of changes in the phenotypes and functions [75]. Furthermore, dystrophic microglia are both found in both the aged and AD brain [76]. Recent research reveals a direct link between inflammasome activation and age-related functional decline [77], meanwhile, neuroinflammation induced by microglia has been proofed to be a vital factor in the progression of AD, suggesting the close relationship among immune-aging, neuroinflammation, and AD.
Neuroinflammation usually refers to a CNS specific, noninfectious chronic inflammation-like glial response that leads to neuronal degradation [78]. Microglia participates in the neuroinflammation through a series of intercellular communications. Reactive microglia release proinflammatory cytokines such as interleukin-1beta (IL-1β), IL-6, IL-18, and tumor necrosis factor-alpha (TNF-α), and up-regulates the expression of chemokines such as CCL2, CCR3, and CCR5, resulting in local inflammatory responses, causing the death of neural cells [79,80,81,82]. In AD, the microglia is able to contribute to the neuroinflammation under the stimulation of Aβ oligomers and plagues via cell surface receptors, including toll-liker receptors (TLRs); LPS receptor cluster of differentiation 14 (CD14); and scavenger receptors such as SCARA1, CD36, and CD47 [83,84,85,86,87]. CD36 is able to recognize and bind with Aβ, causing the activation of the NACHT, LRR, and PYD domains-containing protein 3 (NLRP3) inflammasome, which then activates proinflammatory cytokines IL-1β (Figure 1). Meanwhile, the knockout of CD36 in the macrophages prevents the activation of NLRP3 inflammasome, IL-1β release, and intracellular Aβ accumulation [88]. NLRP3 belongs to the NOD-like receptor superfamily, mainly expressed in the microglia in the brain, functioning as a pattern recognition receptor (PRR) in the innate immune system. NLRP3 inflammasome is activated in AD [89,90]. The deletion of either NLRP3 or downstream regulator caspase-1 is able to rescue the symptoms of AD in APP/SP1 mice [90]. Furthermore, the association of immune receptors, including TREM2 and CD33 with Alzheimer’s disease, indicate the important part neuroinflammation played in this disease [22,91,92,93]. In sporadic AD, reduced Aβ clearance can be linked with insufficient microglial phagocytic capacity, which is characterized by the down-regulation of Aβ phagocytosis receptors and increased cytokine concentrations [94,95].
Microglia activation can be either beneficial or detrimental in the pathogenesis of AD. Research indicates that early activations of microglia in AD present neuroprotective functions by promoting Aβ clearance, however, in response to Aβ aggregation along with the progress of the disease, proinflammatory cytokines production down-regulates the expression of Aβ clearance-related components, and hence, in turn, promotes the Aβ aggregation and neurodegeneration [94].
Peripheral macrophages present a diverse range of phenotypic states, from the proinflammatory M1 phenotype to the alternative activation M2 phenotype, especially under chronic inflammation conditions [96]. Likewise, microglia also presents similar polarizations during chronic inflammation, although the exact classifications are still under debate [97,98]. Generally, the LPS-induced M1-like phenotype of the reactive microglia releases proinflammatory cytokines, such as IL-1β, IL-6, IL-18, TNF-α, and ROS, in order to fight against pathogens invasion and the tumor cell; meanwhile, the M2-like phenotype of microglia produces predominantly anti-inflammatory cytokines, such as IL-10, IL-4, IL-13, and TGF-β [97,98]. In vivo study has revealed that the presence of Aβ is able to induce the phagocytosis of microglia [99], which indicates the protective function of reactive microglia against Aβ accumulation. However, in vitro evidence suggests that the ability of the phagocytosis of microglia is inhibited in AD [100]. LPS induced M1-like reactive microglia present significantly reduced activity in the phagocytosis of Aβ, while this reduction can be rescued by IL-4 induced M2-like microglia activation [100]. Moreover, proinflammatory cytokines TNF-α and IFNγ are able to inhibit Aβ uptake and internalized degradation, while anti-inflammatory cytokines IL-10 promote this ability [101,102]. This phenomenon give rise to a hypothesis that indicates that Aβ accumulation in AD may be due to the alternation of the microglia phenotype. Therefore, the modulation of microglia towards M2-like reactive microglia may have its potential benefit in AD treatment (i.e., the conversion of reactive microglia from a detrimental to beneficial factor). Recent research suggests that the inhibition of NLRP3 inflammasome using a new drug, MCC950, or its downstream effector caspase-1, is able to increase the clearance of Aβ and improve the cognitive function in APP/PS1 mice [103,104]. Together with the evidence indicating that the polymorphisms of NLRP3 are related to the incidence of sporadic AD, this suggests an important role of NLRP3 inflammasome in converting microglia for beneficial effects, and as a promising new target for AD treatments [105,106].

4. Drugs for Alzheimer’s Disease Treatment

4.1. Current Drugs

The drugs currently used for AD can only temporarily relieve its symptoms; meanwhile, no medication is able to stop or reverse the underling progress of this disease. The loss of neurons and synapse in the brain is considered the most direct causes of symptoms in AD. Currently, only five drugs have been approved by the FDA for clinical treatments of AD. Four drugs, tacrine, rivastigmine, galantamine, and donepezil, are acetylcholinesterase inhibitors (AChEIs), thereby enhancing the concentration and duration of the action of the neurotransmitter acetylcholine (Ach). Tacrine is the first approved drug for mild to moderate AD [107]. Because of its safety issue on hepatotoxicity and its uncertain efficacy, it was discontinued in the USA in 2013 [108,109,110,111]. Rivastigmine and galantamine are proved to have a beneficial effect in the treatment of mild to moderate dementia of the Alzheimer’s type [112,113,114,115], while donepezil is used for the treatment of mild, moderate, and severe dementia in AD [116,117]. AChEIs are used to improve the cognition and behavior of AD patients, but are not able to slow down the progression of or reverse the disease; meanwhile, the use of these drugs is usually accompanied with common adverse effects such as nausea, vomiting, diarrhea, headaches, and dizziness [118,119,120]. Memantine, another clinically used drug is an N-Methyl-d-aspartate (NMDA) receptor antagonist. In the brain of AD patients, NMDA receptors are over simulated because of the overload release of glutamate by neurons, causing an increased level of calcium influx and neuronal cell death. Memantine displays some beneficial effects on the symptoms in moderate-to-severe Alzheimer’s disease, whereas it displays limited effectiveness in the mild disease [121,122].

4.2. Drugs Targeting Amyloidogenic Route

As a key component in the pathogenesis of AD, extracellular Aβ aggregation is one of the most studied targets for drug development. Strategies mostly include the prevention of Aβ production and aggregation, as well as anti-Aβ vaccines for immunotherapies. Drugs targeting Aβ is one of the most studied areas in AD. Generally, the strategies for drug development include the reduction of Aβ production or aggregation, and the promotion of Aβ clearance (Table 1). In order to reduce the production of Aβ, three enzymes are usually targeted, namely: α-secretase, β-secretase, and γ-secretase. This can be approached by enhancing the activity of α-secretase or by suppressing the β- and γ-secretase.

4.2.1. α-Secretase Enhancer

Acitretin is a retinoid drug that has been wildly used in the treatment of dermatologic diseases such as psoriasis and hidradenitis suppurativa (HS) [123,124,125], which has recently been reported to act as α-secretase enhancer promoting the non-amyloidogenic APP pathway in patients with mild to moderate AD [126]. Its phase II clinical trials (NCT01078168) have recently been completed, showing a significant increase in the cerebrospinal fluid (CSF) soluble alpha-cleaved amyloid precursor protein (APPsα) concentration in patients receiving oral treatment of acitretin over patients receiving the placebo. Epigallocatechin-gallate (EGCG) is a green tea polyphenol that is suggested to have neuroprotective properties in neurodegenerative diseases [127,128], and has completed phase II and III clinical trials (NCT00951834) for early stage AD. Etazolate (EHT0202) as a γ-aminobutyric acid (GABA)A receptor modulator and α-secretase activator is proven to present neuroprotective functions [129,130]. Its phase II clinical trial in AD treatment (NCT00880412) was completed in 2009, although no results or further studies have been published since then.

4.2.2. β-Secretase Inhibitor

Lanabecestat (also referred as AZD3293 and LY3314814) is a β-secretase inhibitor with previous studies indicating the long time reduction of plasma Aβ in AD patients [131,132]. Multiple phase II and III clinical trials have been conducted for early to mild AD treatment, however they were later terminated because of a lack of effect (NCT02245737, NCT02972658, and NCT02783573) [133]. LY3202626 is a β-secretase inhibitor that was in phase II clinical trials, which were terminated recently because of a small probability of achieving significant results (NCT02791191). LY2886721 was terminated in the phase I/II trials because of abnormal liver biochemical tests in some of the participants (NCT01561430). Verubecestat (MK-8931) is a β-secretase inhibitor reported to reduce CNS Aβ in both animal models and AD patients [134,135,136,137]. However, two clinical trials of this drug were terminated recently (NCT01739348 and NCT01953601). Multiple failures in β-secretase inhibitors for AD treatment led to a change of strategy to target the patients in the early stages of AD, or individuals with a higher risk [133]. Elenbecestat (E2609) is a small molecule β-secretase inhibitor with two phase III clinical trials (NCT02956486 and NCT03036280) currently ongoing for efficacy and safety evaluation in early AD treatment, and one phase II trial (NCT02322021) for MCI and mild to moderate AD. Atabecestat (JNJ-54861911) is a β-secretase inhibitor proven to reduce CSF-Aβ concentrations in a dose-dependent manner in both healthy elderlies and early stage AD patients [138,139,140]. Two phase II/III clinical trials of this drug are still ongoing (NCT01760005 and NCT02569398). However, an extension study for its long-term safety and tolerability was terminated recently because of benefit risk (NCT02406027). CNP520 is a β-secretase inhibitor developed for AD patients in the very early stages [141,142]. Prevention trials of CNP520 in healthy adults over 60 years-old and individuals carrying the APOE ε4 allele show a significant dose-dependent reduction in the CSF-Aβ concentrations [141,142]. Two phase II/III clinical trials of this drug in healthy participants at risk of onset of AD (60–70 years-old and APOE ε4 allele carriers) are ongoing (NCT03131453 and NCT02565511).

4.2.3. γ-Secretase Inhibitor

Gamma-secretase inhibitor is another approach to reduce Aβ production. Semagacestat is the first γ-secretase inhibitor that has entered the phase III clinical trials. It is reported to have dose-dependent cognitive and functional worsening, which led to the termination of its tests. Other adverse effects include weight loss, risk of skin cancer, and infection (NCT01035138, NCT00762411, and NCT00594568) [143]. Another γ-secretase inhibitor, Avagacestat, was discontinued because of a lack of efficacy [143,144,145,146].

4.2.4. Aβ Aggregation Inhibitor

Another approach targeting the amyloidogenic route is inhibiting Aβ aggregation (Table 1). PBT2 is a metal protein-attenuating compound (MPAC), however its phase II/III clinical trial was announced as failed because of a lack of efficacy [143,147]. Scyllo-inositol is an inositol stereoisomer and Aβ aggregation inhibitor. Its phase II clinical trial demonstrated some beneficial effects in mild to moderate AD, however was insufficient for a robust benefit conclusion because of its small sample size [148,149]. Tramiprosate (homotaurine, 3APS) is an amino acid originally found in seaweed, with homology with taurine and 4-aminobutyrate (γ-aminobutyric acid—GABA), act as a GABA receptor agonist, and therefore demonstrates neuroprotective effects [150]. Most importantly, tramiprosate is able to inhibit Aβ aggregation by binding to Aβ and preventing β-sheet formation [150]. In vitro and in vivo studies have revealed that tramiprosate is able to attenuate the long-term potentiation (LTP) inhibition cause by Aβ toxicity, and dose-dependently reduce soluble and insoluble Aβ in transgenic mice [151,152]. However, its phase III clinical trials have been marked as unknown, with the last update date back in 2007 (NCT00314912, NCT00217763, and NCT00088673). Later, in 2009, a publication revealed that tramiprosate treatments in mild to moderate AD patients showed some beneficial effects, but did not reach statistical significance [153]. Recent studies of tramiprosate demonstrated the clinical benefits of this drug in ApoE ε4/ε4 patients [154,155,156,157]. New drugs developed based on tramiprosate—including ALZ-801, a prodrug of tramiprosate, and GQD-T, the combination of graphene quantum dots (GQDs) and tramiprosate—have both demonstrated promising results in AD treatment [158,159,160]. Sodium oligo-mannurarate (GV-971) is an Aβ aggregation inhibitor [161], and recently completed a 36-week phase III clinical trial for the treatment of mild to moderate AD (NCT02293915). A new phase I clinical trial of GV-971 is currently investigating the safety and pharmacokinetics studies (NCT03715114)

4.2.5. Aβ Vaccines

Immunotherapies targeting Aβ include vaccines and antibodies. A first generation vaccine, AN-1792, was terminated because of the incidence of adverse effects such as cerebral inflammation (NCT00021723) [162]. A follow-up study of the long term effect of this drug in AD revealed a significantly lower Aβ volume in the brain of AD patients, however no significant improvement in the symptoms of dementia [163]. Followed by this, new vaccines, CAD 106 and Vanutide cridificar (ACC-001), recently finish phase II clinical trials, with promising results [143,164,165,166]. A phase II/III clinical trial of CAD106 is still ongoing (NCT02565511).

4.2.6. Aβ Antibodies

Anti-Aβ antibodies are termed as passive immunotherapy. Bapineuzumab and solanezumab are monoclonal antibodies against Aβ (1–6) and Aβ (12–28), respectively [167,168]. In 2012, a clinical trial of bapineuzumab was terminated at phase III. Its treatment in AD patients displayed a significant decrease in the senile plaques and tau protein in CSF, meanwhile, showed no improvement in the cognitive functions [143]. Three phase III trials of solanezumab were recently terminated because of a lack of efficiency in prodromal and mild AD (NCT01127633, NCT01900665, and NCT02760602). Two other phase II/III clinical trials for old individuals with at risk of memory loss and familial AD are currently active (NCT02008357 and NCT01760005). Ponezumab is a human monoclonal antibody against Aβ40. Ponezumab administration demonstrates a dose-dependent increase in plasma Aβ40 and a decrease in the hippocampal Aβ concentration in transgenic mice [169]. Its phase II trials were completed with results indicating a dose-dependentl increase of plasma Aβ, with no effect in the CSF biomarkers, brain Aβ load, and cognitive improvements [170,171,172]. GSK933776 produced promising results in phase I clinical trials [173,174], however no new clinical trial has been conducted since 2013. LY2599666 is an Aβ antibody for MCI or AD, and was terminated in a phase I trial because of a lack of efficacy (NCT02614131). Octagam® 10% (IVIG) is an immune globulin intravenous (10% solution). Two phase III clinical trials of Octagam® 10% for AD were terminated because of a lack of efficacy (NCT01736579 and NCT01524887), while other phase II/III trials for AD and mild cognitive impairment (MCI) are currently active (NCT01561053, NCT01300728, and NCT03319810). Aducanumab (BIIB037) is a human monoclonal antibody against aggregated forms of Aβ, including soluble Aβ oligomer and insoluble Aβ fibrils [175,176]. Aducanumab treatments have demonstrated beneficial effects in Aβ plaque clearing and a reverse calcium dysfunction in a Tg2576 mice model of AD [177,178]. In an ongoing phase I clinical trial, aducanumab treatment is able to reduce Aβ plaques and attenuate the decline of cognitive functions of prodromal or mild AD patients (NCT01677572) [179,180,181]. Two phase III clinical trials of aducanumab in early AD, and two trials for MCI or mild AD are currently ongoing (NCT02484547, NCT02477800, NCT03639987, and NCT01677572). Crenezumab is a monoclonal IgG4 Aβ antibody able to recognize multiple forms of Aβ, especially oligomers, and inhibit Aβ aggregation [182]. Although previous clinical trials of this drug failed to significantly improve the symptoms of mild or moderate AD, some beneficial effects with high dose crenezumab indicated potential treatment effects [183,184]. Multiple clinical trials of crenezumab are currently ongoing (NCT03491150, NCT03114657, NCT02670083, NCT01998841, and NCT02353598). Gantenerumab is a fully human IgG1 Aβ antibody that recognizes a conformational epitope of Aβ fibrils in the sub-nanomolar affinity in vitro, and therefore can bind to the aggregated Aβ with a high affinity [185]. The experimental results demonstrate that gantenerumab is able to enhance the phagocytosis of Aβ in brain slices co-cultured with macrophages, neutralize the inhibitory effects on LTP caused by Aβ42 oligomers in rat brains, promote cerebral Aβ clearance by recruiting microglia in a transgenic mouse model of AD, and therefore inhibit the formation of Aβ plaque [185]. Clinical trials of this drug have come out showing a reduction in the cerebral Aβ in AD patients [186,187,188]. Recently, a phase III clinical trial of gantenerumab in prodromal AD was stopped because of a lack of effects, however some dose-dependent beneficial effects indicated a probability of reaching significance with higher doses [189]. Four phase III clinical trials of gantenerumab are active for prodromal to mild AD (NCT03443973, NCT03444870, NCT02051608, and NCT01224106). The monoclonal antibody for Aβ fibrillars SAR228810 is a next generation product of murine antibody SAR255952, and has completed a phase I clinical trial with no further clinical studies ongoing.

4.3. Drugs Targeting Tau Protein

Currently, several drugs targeting tau protein have entered clinical trials (Table 1). Aggregation inhibitors, TRx0014 and LMTM (TRx0237), are methylene blue dye derivatives that are able to prevent tau and amyloid aggregation [190]. LMTM is the next generation drug of TRx0014, which has produced promising results in clinical trials [143]. However, in recent trials, LMTM demonstrated no significant beneficial effect in the add-on therapy of mild to moderate AD (NCT01689246) [191]. Meanwhile, another phase III clinical trial (NCT01689233) for LMTM as a monotherapy for mild to moderate AD demonstrated promising results, indicating that LMTM might be effective in future studies [192]. New phase II/III clinical trials of LMTM for early AD (NCT03446001) and mild to moderate AD (NCT03539380) have recently started. Tideglusib is a GSK3-β inhibitor able to prevent tau hyperphosphorylation, however phase II clinical trials of this drug display no significant efficacy to AD [193].
Human tau antibody ABBV-8E12 is safe to use and is currently in phase II clinical trials for early AD (NCT02880956 and NCT03712787) [194]. RO7105705 is an antibody targeting extracellular tau, therefore stopping the spread of pathological tau, demonstrating effective outcomes in preventing tau pathology in tau-P301 transgenic mice, and attenuating microglia induced inflammation [195]. One phase II clinical trial of RO7105705 for prodromal to mild AD is ongoing (NCT03289143). AADvac1 is the first human pathologically modified tau vaccine for active immunotherapy, able to reduce 95% of the hyperphosphorylation of tau and improve the symptoms of transgenic mice [196]. Currently, early phase clinical trials have been completed with promising results [197,198]; a phase II clinical trial is ongoing for mild AD (NCT02579252). TPI 287 is a microtubule stabilizer that able to bind on tubulin and stabilize the microtubule [199], and is currently in a phase I clinical trial (NCT01966666).

4.4. Drugs Targeting Inflammation

4.4.1. Non-Steroidal Anti-Inflammatory Drugs

The potential beneficial effect in the non-steroidal anti-inflammatory drugs (NSAIDs) has drawn public attention because several systematic reviews revealed that long term users of NSAIDs showed lower risks in the incidence of AD [200,201,202,203]. In a recent study, the NSAIDs of the fenamate class displayed a significant effect in inhibiting the NLRP3 inflammasome activation in vitro, and attenuated microglia activation in transgenic mice [204]. However, randomized controlled trials (RCT) for relationships of NSAIDs usage and AD risk failed to show significance among the normal population without dementia or AD patients [205,206,207], leaving the effectiveness of NSAIDs in AD treatment controversial.
Ibuprofen is one of the most used NSAIDs reported to have a protective effect against incidence of AD, along with other NSAIDs [202]; later, RCTs proved it to have no effect on cognitive progresses in AD [207,208]. One clinical trial of ibuprofen in AD was marked as unknown (NCT00239746), with no update since 2009. A recent phase III clinical trial of a combination treatment of ibuprofen and cromolyn (ALZT-OP1) in early AD is ongoing (NCT02547818). Tarenflurbil is a NSAID structurally related to ibuprofen, but its poor ability in brain penetration and its lack of efficacy for AD treatment has led to failure in phase III clinical trials [143,209]. Salsalate is a NSAID that is currently in an ongoing phase I clinical trial in patients with mild to moderate AD (NCT03277573). Celecoxib is also a NSAID whose phase III clinical trials for AD treatment were completed in 2016 (NCT00007189)

4.4.2. Other Drugs Targeting Inflammation

Resveratrol is an antioxidant that recently completed a phase II clinical trial for the treatment of AD. As a result, resveratrol was found to be safe and well-tolerated, with the ability to cross the blood–brain barrier, although further study is required in order to determine its efficacy in AD treatment [210]. Etanercept is a TNF-α inhibitor. The phase II study of its safety and tolerability of this drug was conducted recently, and showed positive results [211]. Simvastatin is a cholesterol targeting 3-hydroxy-3-methyl-glutaryl-coenzyme A reductase (HMG-CoA reductase) inhibitor. Previous studies indicated that despite a significant decrease in the cholesterol level, no marked beneficial effect is associated with the use of this drug [212]. Neflamapimod (VX-745) is a selective inhibitor of the α isoform of the mitogen-activated serine/threonine protein kinase p38 MAPK (p38 MAPKα), and is reported to be able to slow the progression of transgenic AD mice [213]. A phase II clinical trial of neflamapimod demonstrated improvement in the episodic memory of AD patients [214]. Two phase II clinical trials of this drug in the inflammation in AD are currently ongoing (NCT03435861 and NCT03402659). Azeliragon (TTP488) is a small molecule inhibitor of the receptor for advanced glycation endproducts (RAGE) and has been reported to slow cognitive decline in AD patients [215,216], however two phase III clinical trials were terminated recently because of a lack of efficacy (NCT02080364 and NCT02916056). Pioglitazone is a peroxisome-proliferator activated receptor γ (PPARγ) agonist that showed promising results in AD treatment, however two phase III trials were terminated recently because of a lack of efficacy (NCT01931566 and NCT02284906) [217,218,219].

5. Conclusive Remarks

Alzheimer’s disease has become an important disease affecting the healthy aging of humans. AD is caused by multiple factors and cannot be explained by a solo hypothesis. Currently, 2103 studies can be found on the clinicaltrials.gov website under the category of Alzheimer’s disease, including drugs, therapies, and imaging markers. Drugs targeting different aspects of this disease have been intensively studied, however no effective drug has been developed for clinical use; meanwhile, many promising drug developments have been terminated at late clinical trials. For the recent decades, Aβ and tau accumulation in the brain, as the most used hallmarks of AD, have been intensively studied as drug targets for a promising cure for this disease. However, to date, drugs targeting Aβ or tau are able to present only limited beneficial effects on the pathogenesis of AD, which results in the concern of strategies for adjustment in drug development in AD. Recent advances in the understanding of the important role played by chronic neuroinflammation induced by microglia in AD, indicates a potential target for AD treatment.

Author Contributions

Y.D. wrote the manuscript and prepared the figure; X.L. reviewed the manuscript; L.H. and J.C. conceived the review topic and performed a comprehensive review of the literature.

Funding

This work was supported by grants from the China Postdoctoral Science Foundation (grant no. 2017M622142) and the Special Fund for Post-Doctoral Innovative Projects in the Shandong Province of China (grant no. 201703029).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Burns, A.; Iliffe, S. Alzheimer’s disease. BMJ 2009, 338, b158. [Google Scholar] [CrossRef] [PubMed]
  2. Mendez, M.F. Early-onset Alzheimer’s disease: Nonamnestic subtypes and type 2 AD. Arch. Med. Res. 2012, 43, 677–685. [Google Scholar] [CrossRef] [PubMed]
  3. Alzheimer’s Association. 2016 Alzheimer’s disease facts and figures. Alzheimer’s Dement. J. Alzheimer’s Assoc. 2016, 12, 459–509. [Google Scholar] [CrossRef]
  4. Mielke, M.M.; Vemuri, P.; Rocca, W.A. Clinical epidemiology of Alzheimer’s disease: Assessing sex and gender differences. Clin. Epidemiol. 2014, 6, 37–48. [Google Scholar] [CrossRef] [PubMed]
  5. Hebert, L.E.; Weuve, J.; Scherr, P.A.; Evans, D.A. Alzheimer disease in the United States (2010-2050) estimated using the 2010 census. Neurology 2013, 80, 1778–1783. [Google Scholar] [CrossRef] [Green Version]
  6. Tom, S.E.; Hubbard, R.A.; Crane, P.K.; Haneuse, S.J.; Bowen, J.; McCormick, W.C.; McCurry, S.; Larson, E.B. Characterization of dementia and Alzheimer’s disease in an older population: Updated incidence and life expectancy with and without dementia. Am. J. Public Health 2015, 105, 408–413. [Google Scholar] [CrossRef]
  7. Bonomo, S.M.; Rigamonti, A.E.; Giunta, M.; Galimberti, D.; Guaita, A.; Gagliano, M.G.; Muller, E.E.; Cella, S.G. Menopausal transition: A possible risk factor for brain pathologic events. Neurobiol. Aging 2009, 30, 71–80. [Google Scholar] [CrossRef]
  8. Pike, C.J.; Carroll, J.C.; Rosario, E.R.; Barron, A.M. Protective actions of sex steroid hormones in Alzheimer’s disease. Front. Neuroendocrinol. 2009, 30, 239–258. [Google Scholar] [CrossRef] [Green Version]
  9. Srivastava, D.P.; Woolfrey, K.M.; Penzes, P. Insights into rapid modulation of neuroplasticity by brain estrogens. Pharmacol. Rev. 2013, 65, 1318–1350. [Google Scholar] [CrossRef]
  10. Stern, Y. Cognitive reserve in ageing and Alzheimer’s disease. Lancet Neurol. 2012, 11, 1006–1012. [Google Scholar] [CrossRef]
  11. Rocca, W.A.; Mielke, M.M.; Vemuri, P.; Miller, V.M. Sex and gender differences in the causes of dementia: A narrative review. Maturitas 2014, 79, 196–201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Berchtold, N.C.; Cotman, C.W. Evolution in the conceptualization of dementia and Alzheimer’s disease: Greco-Roman period to the 1960s. Neurobiol. Aging 1998, 19, 173–189. [Google Scholar] [CrossRef]
  13. Zhang, B.; Gaiteri, C.; Bodea, L.G.; Wang, Z.; McElwee, J.; Podtelezhnikov, A.A.; Zhang, C.; Xie, T.; Tran, L.; Dobrin, R.; et al. Integrated systems approach identifies genetic nodes and networks in late-onset Alzheimer’s disease. Cell 2013, 153, 707–720. [Google Scholar] [CrossRef]
  14. Carreiras, M.C.; Mendes, E.; Perry, M.J.; Francisco, A.P.; Marco-Contelles, J. The multifactorial nature of Alzheimer’s disease for developing potential therapeutics. Curr. Top. Med. Chem. 2013, 13, 1745–1770. [Google Scholar] [CrossRef] [PubMed]
  15. Arendt, T. Synaptic degeneration in Alzheimer’s disease. Acta Neuropathol. 2009, 118, 167–179. [Google Scholar] [CrossRef] [PubMed]
  16. Blennow, K.; de Leon, M.J.; Zetterberg, H. Alzheimer’s disease. Lancet 2006, 368, 387–403. [Google Scholar] [CrossRef]
  17. Campion, D.; Dumanchin, C.; Hannequin, D.; Dubois, B.; Belliard, S.; Puel, M.; Thomas-Anterion, C.; Michon, A.; Martin, C.; Charbonnier, F.; et al. Early-onset autosomal dominant Alzheimer disease: Prevalence, genetic heterogeneity, and mutation spectrum. Am. J. Hum. Genet. 1999, 65, 664–670. [Google Scholar] [CrossRef]
  18. Nussbaum, R.L. Genome-wide association studies, Alzheimer disease, and understudied populations. JAMA 2013, 309, 1527–1528. [Google Scholar] [CrossRef]
  19. Mahley, R.W.; Weisgraber, K.H.; Huang, Y. Apolipoprotein E4: A causative factor and therapeutic target in neuropathology, including Alzheimer’s disease. Proc. Natl. Acad. Sci. USA 2006, 103, 5644–5651. [Google Scholar] [CrossRef]
  20. Neu, S.C.; Pa, J.; Kukull, W.; Beekly, D.; Kuzma, A.; Gangadharan, P.; Wang, L.S.; Romero, K.; Arneric, S.P.; Redolfi, A.; et al. Apolipoprotein E Genotype and Sex Risk Factors for Alzheimer Disease: A Meta-analysis. JAMA Neurol. 2017, 74, 1178–1189. [Google Scholar] [CrossRef]
  21. Jonsson, T.; Stefansson, H.; Steinberg, S.; Jonsdottir, I.; Jonsson, P.V.; Snaedal, J.; Bjornsson, S.; Huttenlocher, J.; Levey, A.I.; Lah, J.J.; et al. Variant of TREM2 associated with the risk of Alzheimer’s disease. N. Engl. J. Med. 2013, 368, 107–116. [Google Scholar] [CrossRef] [PubMed]
  22. Guerreiro, R.; Wojtas, A.; Bras, J.; Carrasquillo, M.; Rogaeva, E.; Majounie, E.; Cruchaga, C.; Sassi, C.; Kauwe, J.S.; Younkin, S.; et al. TREM2 variants in Alzheimer’s disease. N. Engl. J. Med. 2013, 368, 117–127. [Google Scholar] [CrossRef] [PubMed]
  23. Abduljaleel, Z.; Al-Allaf, F.A.; Khan, W.; Athar, M.; Shahzad, N.; Taher, M.M.; Elrobh, M.; Alanazi, M.S.; El-Huneidi, W. Evidence of trem2 variant associated with triple risk of Alzheimer’s disease. PLoS ONE 2014, 9, e92648. [Google Scholar] [CrossRef] [PubMed]
  24. Jin, S.C.; Benitez, B.A.; Karch, C.M.; Cooper, B.; Skorupa, T.; Carrell, D.; Norton, J.B.; Hsu, S.; Harari, O.; Cai, Y.; et al. Coding variants in TREM2 increase risk for Alzheimer’s disease. Hum. Mol. Genet. 2014, 23, 5838–5846. [Google Scholar] [CrossRef] [PubMed]
  25. Lu, Y.; Liu, W.; Wang, X. TREM2 variants and risk of Alzheimer’s disease: A meta-analysis. Neurol. Sci. 2015, 36, 1881–1888. [Google Scholar] [CrossRef] [PubMed]
  26. Bezprozvanny, I.; Mattson, M.P. Neuronal calcium mishandling and the pathogenesis of Alzheimer’s disease. Trends Neurosci. 2008, 31, 454–463. [Google Scholar] [CrossRef] [PubMed]
  27. Bezprozvanny, I. Calcium signaling and neurodegenerative diseases. Trends Mol. Med. 2009, 15, 89–100. [Google Scholar] [CrossRef] [Green Version]
  28. Berridge, M.J. Dysregulation of neural calcium signaling in Alzheimer disease, bipolar disorder and schizophrenia. Prion 2013, 7, 2–13. [Google Scholar] [CrossRef] [Green Version]
  29. Berridge, M.J.; Bootman, M.D.; Roderick, H.L. Calcium signalling: Dynamics, homeostasis and remodelling. Nature Rev. Mol. Cell Biol. 2003, 4, 517–529. [Google Scholar] [CrossRef]
  30. Magi, S.; Castaldo, P.; Macri, M.L.; Maiolino, M.; Matteucci, A.; Bastioli, G.; Gratteri, S.; Amoroso, S.; Lariccia, V. Intracellular Calcium Dysregulation: Implications for Alzheimer’s Disease. BioMed Res. Int. 2016, 2016, 6701324. [Google Scholar] [CrossRef]
  31. Dreses-Werringloer, U.; Vingtdeux, V.; Zhao, H.; Chandakkar, P.; Davies, P.; Marambaud, P. CALHM1 controls the Ca(2)(+)-dependent MEK, ERK, RSK and MSK signaling cascade in neurons. J. Cell Sci. 2013, 126, 1199–1206. [Google Scholar] [CrossRef] [PubMed]
  32. Vingtdeux, V.; Chang, E.H.; Frattini, S.A.; Zhao, H.; Chandakkar, P.; Adrien, L.; Strohl, J.J.; Gibson, E.L.; Ohmoto, M.; Matsumoto, I.; et al. CALHM1 deficiency impairs cerebral neuron activity and memory flexibility in mice. Sci. Rep. 2016, 6, 24250. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Su, B.; Wang, X.; Nunomura, A.; Moreira, P.I.; Lee, H.G.; Perry, G.; Smith, M.A.; Zhu, X. Oxidative stress signaling in Alzheimer’s disease. Curr. Alzheimer Res. 2008, 5, 525–532. [Google Scholar] [CrossRef] [PubMed]
  34. Onyango, I.G.; Khan, S.M. Oxidative stress, mitochondrial dysfunction, and stress signaling in Alzheimer’s disease. Curr. Alzheimer Res. 2006, 3, 339–349. [Google Scholar] [CrossRef] [PubMed]
  35. Liao, Y.; Dong, Y.; Cheng, J. The Function of the Mitochondrial Calcium Uniporter in Neurodegenerative Disorders. Int. J. Mol. Sci. 2017, 18. [Google Scholar] [CrossRef] [PubMed]
  36. Cheng, J.; Liao, Y.; Zhou, L.; Peng, S.; Chen, H.; Yuan, Z. Amplified RLR signaling activation through an interferon-stimulated gene-endoplasmic reticulum stress-mitochondrial calcium uniporter protein loop. Sci. Rep. 2016, 6, 20158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Liao, Y.; Hao, Y.; Chen, H.; He, Q.; Yuan, Z.; Cheng, J. Mitochondrial calcium uniporter protein MCU is involved in oxidative stress-induced cell death. Protein Cell 2015, 6, 434–442. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Lambert, J.C.; Ibrahim-Verbaas, C.A.; Harold, D.; Naj, A.C.; Sims, R.; Bellenguez, C.; DeStafano, A.L.; Bis, J.C.; Beecham, G.W.; Grenier-Boley, B.; et al. Meta-analysis of 74,046 individuals identifies 11 new susceptibility loci for Alzheimer’s disease. Nat. Genet. 2013, 45, 1452–1458. [Google Scholar] [CrossRef] [PubMed]
  39. Francis, P.T.; Palmer, A.M.; Snape, M.; Wilcock, G.K. The cholinergic hypothesis of Alzheimer’s disease: A review of progress. J. Neurol. Neurosurg. Psychiatry 1999, 66, 137–147. [Google Scholar] [CrossRef] [PubMed]
  40. Martorana, A.; Esposito, Z.; Koch, G. Beyond the cholinergic hypothesis: Do current drugs work in Alzheimer’s disease? CNS Neurosci. Ther. 2010, 16, 235–245. [Google Scholar] [CrossRef] [PubMed]
  41. Barage, S.H.; Sonawane, K.D. Amyloid cascade hypothesis: Pathogenesis and therapeutic strategies in Alzheimer’s disease. Neuropeptides 2015, 52, 1–18. [Google Scholar] [CrossRef] [PubMed]
  42. Karran, E.; Mercken, M.; De Strooper, B. The amyloid cascade hypothesis for Alzheimer’s disease: An appraisal for the development of therapeutics. Nature Rev. Drug Discov. 2011, 10, 698–712. [Google Scholar] [CrossRef] [PubMed]
  43. Selkoe, D.J. Alzheimer’s disease: Genes, proteins, and therapy. Physiol. Rev. 2001, 81, 741–766. [Google Scholar] [CrossRef] [PubMed]
  44. Haass, C. Take five–BACE and the gamma-secretase quartet conduct Alzheimer’s amyloid beta-peptide generation. EMBO J. 2004, 23, 483–488. [Google Scholar] [CrossRef] [PubMed]
  45. De Strooper, B.; Vassar, R.; Golde, T. The secretases: Enzymes with therapeutic potential in Alzheimer disease. Nat. Rev. Neurol. 2010, 6, 99–107. [Google Scholar] [CrossRef] [PubMed]
  46. 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] [PubMed]
  47. Lalonde, R.; Dumont, M.; Staufenbiel, M.; Sturchler-Pierrat, C.; Strazielle, C. Spatial learning, exploration, anxiety, and motor coordination in female APP23 transgenic mice with the Swedish mutation. Brain Res. 2002, 956, 36–44. [Google Scholar] [CrossRef]
  48. Jiang, Q.; Lee, C.Y.; Mandrekar, S.; Wilkinson, B.; Cramer, P.; Zelcer, N.; Mann, K.; Lamb, B.; Willson, T.M.; Collins, J.L.; et al. ApoE promotes the proteolytic degradation of Abeta. Neuron 2008, 58, 681–693. [Google Scholar] [CrossRef]
  49. Hauser, P.S.; Ryan, R.O. Impact of apolipoprotein E on Alzheimer’s disease. Curr. Alzheimer Res. 2013, 10, 809–817. [Google Scholar] [CrossRef]
  50. Liu, J.; Chang, L.; Roselli, F.; Almeida, O.F.; Gao, X.; Wang, X.; Yew, D.T.; Wu, Y. Amyloid-beta induces caspase-dependent loss of PSD-95 and synaptophysin through NMDA receptors. J. Alzheimer’s Dis. JAD 2010, 22, 541–556. [Google Scholar] [CrossRef]
  51. Roberson, E.D.; Scearce-Levie, K.; Palop, J.J.; Yan, F.; Cheng, I.H.; Wu, T.; Gerstein, H.; Yu, G.Q.; Mucke, L. Reducing endogenous tau ameliorates amyloid beta-induced deficits in an Alzheimer’s disease mouse model. Science 2007, 316, 750–754. [Google Scholar] [CrossRef] [PubMed]
  52. Nussbaum, J.M.; Seward, M.E.; Bloom, G.S. Alzheimer disease: A tale of two prions. Prion 2013, 7, 14–19. [Google Scholar] [CrossRef]
  53. Borchelt, D.R.; Thinakaran, G.; Eckman, C.B.; Lee, M.K.; Davenport, F.; Ratovitsky, T.; Prada, C.M.; Kim, G.; Seekins, S.; Yager, D.; et al. Familial Alzheimer’s disease-linked presenilin 1 variants elevate Abeta1-42/1-40 ratio in vitro and in vivo. Neuron 1996, 17, 1005–1013. [Google Scholar] [CrossRef]
  54. Cavallucci, V.; D’Amelio, M.; Cecconi, F. Abeta toxicity in Alzheimer’s disease. Mol. Neurobiol. 2012, 45, 366–378. [Google Scholar] [CrossRef] [PubMed]
  55. Lacor, P.N.; Buniel, M.C.; Furlow, P.W.; Clemente, A.S.; Velasco, P.T.; Wood, M.; Viola, K.L.; Klein, W.L. Abeta oligomer-induced aberrations in synapse composition, shape, and density provide a molecular basis for loss of connectivity in Alzheimer’s disease. J. Neurosci. 2007, 27, 796–807. [Google Scholar] [CrossRef]
  56. Nikolaev, A.; McLaughlin, T.; O’Leary, D.D.; Tessier-Lavigne, M. APP binds DR6 to trigger axon pruning and neuron death via distinct caspases. Nature 2009, 457, 981–989. [Google Scholar] [CrossRef] [Green Version]
  57. Hardy, J.A.; Higgins, G.A. Alzheimer’s disease: The amyloid cascade hypothesis. Science 1992, 256, 184–185. [Google Scholar] [CrossRef]
  58. Korczyn, A.D. The amyloid cascade hypothesis. Alzheimer’s Dement. J. Alzheimer’s Assoc. 2008, 4, 176–178. [Google Scholar] [CrossRef]
  59. Ricciarelli, R.; Fedele, E. The amyloid cascade hypothesis in Alzheimer’s disease: It’s time to change our mind. Curr. Neuropharmacol. 2017, 15, 926–935. [Google Scholar] [CrossRef]
  60. Dong, S.; Duan, Y.; Hu, Y.; Zhao, Z. Advances in the pathogenesis of Alzheimer’s disease: A re-evaluation of amyloid cascade hypothesis. Transl. Neurodegener. 2012, 1, 18. [Google Scholar] [CrossRef]
  61. The amyloid cascade hypothesis has misled the pharmaceutical industry. Biochem. Soc. Trans. 2011, 39, 920–923. [CrossRef] [PubMed] [Green Version]
  62. Goedert, M.; Spillantini, M.G.; Crowther, R.A. Tau proteins and neurofibrillary degeneration. Brain Pathol. 1991, 1, 279–286. [Google Scholar] [CrossRef] [PubMed]
  63. Maccioni, R.B.; Munoz, J.P.; Barbeito, L. The molecular bases of Alzheimer’s disease and other neurodegenerative disorders. Arch. Med. Res. 2001, 32, 367–381. [Google Scholar] [CrossRef]
  64. Iqbal, K.; Alonso Adel, C.; Chen, S.; Chohan, M.O.; El-Akkad, E.; Gong, C.X.; Khatoon, S.; Li, B.; Liu, F.; Rahman, A.; et al. Tau pathology in Alzheimer disease and other tauopathies. Biochim. Biophys. Acta 2005, 1739, 198–210. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Chun, W.; Johnson, G.V. The role of tau phosphorylation and cleavage in neuronal cell death. Front. Biosci. J. Virtual Libr. 2007, 12, 733–756. [Google Scholar] [CrossRef]
  66. Cardona, A.E.; Huang, D.; Sasse, M.E.; Ransohoff, R.M. Isolation of murine microglial cells for RNA analysis or flow cytometry. Nat. Protoc. 2006, 1, 1947–1951. [Google Scholar] [CrossRef] [PubMed]
  67. Filiano, A.J.; Gadani, S.P.; Kipnis, J. Interactions of innate and adaptive immunity in brain development and function. Brain Res. 2015, 1617, 18–27. [Google Scholar] [CrossRef] [Green Version]
  68. Wirenfeldt, M.; Babcock, A.A.; Vinters, H.V. Microglia—Insights into immune system structure, function, and reactivity in the central nervous system. Histol. Histopathol. 2011, 26, 519–530. [Google Scholar] [CrossRef]
  69. Aloisi, F. Immune function of microglia. Glia 2001, 36, 165–179. [Google Scholar] [CrossRef]
  70. Kettenmann, H.; Hanisch, U.K.; Noda, M.; Verkhratsky, A. Physiology of microglia. Physiol. Rev. 2011, 91, 461–553. [Google Scholar] [CrossRef]
  71. Ji, K.; Akgul, G.; Wollmuth, L.P.; Tsirka, S.E. Microglia actively regulate the number of functional synapses. PLoS ONE 2013, 8, e56293. [Google Scholar] [CrossRef] [PubMed]
  72. Parkhurst, C.N.; Yang, G.; Ninan, I.; Savas, J.N.; Yates, J.R., 3rd; Lafaille, J.J.; Hempstead, B.L.; Littman, D.R.; Gan, W.B. Microglia promote learning-dependent synapse formation through brain-derived neurotrophic factor. Cell 2013, 155, 1596–1609. [Google Scholar] [CrossRef] [PubMed]
  73. Keren-Shaul, H.; Spinrad, A.; Weiner, A.; Matcovitch-Natan, O.; Dvir-Szternfeld, R.; Ulland, T.K.; David, E.; Baruch, K.; Lara-Astaiso, D.; Toth, B.; et al. A Unique Microglia Type Associated with Restricting Development of Alzheimer’s Disease. Cell 2017, 169, 1276–1290. [Google Scholar] [CrossRef] [PubMed]
  74. Sierra, A.; Gottfried-Blackmore, A.C.; McEwen, B.S.; Bulloch, K. Microglia derived from aging mice exhibit an altered inflammatory profile. Glia 2007, 55, 412–424. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Wong, W.T. Microglial aging in the healthy CNS: Phenotypes, drivers, and rejuvenation. Front. Cell. Neurosci. 2013, 7, 22. [Google Scholar] [CrossRef] [PubMed]
  76. Lopes, K.O.; Sparks, D.L.; Streit, W.J. Microglial dystrophy in the aged and Alzheimer’s disease brain is associated with ferritin immunoreactivity. Glia 2008, 56, 1048–1060. [Google Scholar] [CrossRef] [PubMed]
  77. Youm, Y.H.; Grant, R.W.; McCabe, L.R.; Albarado, D.C.; Nguyen, K.Y.; Ravussin, A.; Pistell, P.; Newman, S.; Carter, R.; Laque, A.; et al. Canonical Nlrp3 inflammasome links systemic low-grade inflammation to functional decline in aging. Cell Metabol. 2013, 18, 519–532. [Google Scholar] [CrossRef]
  78. Streit, W.J. Microglial senescence: Does the brain’s immune system have an expiration date? Trends Neurosci. 2006, 29, 506–510. [Google Scholar] [CrossRef]
  79. Heneka, M.T.; Carson, M.J.; El Khoury, J.; Landreth, G.E.; Brosseron, F.; Feinstein, D.L.; Jacobs, A.H.; Wyss-Coray, T.; Vitorica, J.; Ransohoff, R.M.; et al. Neuroinflammation in Alzheimer’s disease. Lancet Neurol. 2015, 14, 388–405. [Google Scholar] [CrossRef]
  80. Patel, N.S.; Paris, D.; Mathura, V.; Quadros, A.N.; Crawford, F.C.; Mullan, M.J. Inflammatory cytokine levels correlate with amyloid load in transgenic mouse models of Alzheimer’s disease. J. Neuroinflammation 2005, 2, 9. [Google Scholar] [CrossRef]
  81. Xia, M.Q.; Qin, S.X.; Wu, L.J.; Mackay, C.R.; Hyman, B.T. Immunohistochemical study of the beta-chemokine receptors CCR3 and CCR5 and their ligands in normal and Alzheimer’s disease brains. Am. J. Pathol. 1998, 153, 31–37. [Google Scholar] [CrossRef]
  82. Ishizuka, K.; Kimura, T.; Igata-yi, R.; Katsuragi, S.; Takamatsu, J.; Miyakawa, T. Identification of monocyte chemoattractant protein-1 in senile plaques and reactive microglia of Alzheimer’s disease. Psychiatry Clin. Neurosci. 1997, 51, 135–138. [Google Scholar] [CrossRef] [PubMed]
  83. Stewart, C.R.; Stuart, L.M.; Wilkinson, K.; van Gils, J.M.; Deng, J.; Halle, A.; Rayner, K.J.; Boyer, L.; Zhong, R.; Frazier, W.A.; et al. CD36 ligands promote sterile inflammation through assembly of a Toll-like receptor 4 and 6 heterodimer. Nature Immunol. 2010, 11, 155–161. [Google Scholar] [CrossRef] [PubMed]
  84. Bamberger, M.E.; Harris, M.E.; McDonald, D.R.; Husemann, J.; Landreth, G.E. A cell surface receptor complex for fibrillar beta-amyloid mediates microglial activation. J. Neurosci. 2003, 23, 2665–2674. [Google Scholar] [CrossRef] [PubMed]
  85. Fassbender, K.; Walter, S.; Kuhl, S.; Landmann, R.; Ishii, K.; Bertsch, T.; Stalder, A.K.; Muehlhauser, F.; Liu, Y.; Ulmer, A.J.; et al. The LPS receptor (CD14) links innate immunity with Alzheimer’s disease. FASEB J. 2004, 18, 203–205. [Google Scholar] [CrossRef] [PubMed]
  86. El Khoury, J.B.; Moore, K.J.; Means, T.K.; Leung, J.; Terada, K.; Toft, M.; Freeman, M.W.; Luster, A.D. CD36 mediates the innate host response to beta-amyloid. J. Exp. Med. 2003, 197, 1657–1666. [Google Scholar] [CrossRef] [PubMed]
  87. Kagan, J.C.; Horng, T. NLRP3 inflammasome activation: CD36 serves double duty. Nat. Immunol. 2013, 14, 772–774. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Sheedy, F.J.; Grebe, A.; Rayner, K.J.; Kalantari, P.; Ramkhelawon, B.; Carpenter, S.B.; Becker, C.E.; Ediriweera, H.N.; Mullick, A.E.; Golenbock, D.T.; et al. CD36 coordinates NLRP3 inflammasome activation by facilitating intracellular nucleation of soluble ligands into particulate ligands in sterile inflammation. Nat. Immunol. 2013, 14, 812–820. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Saresella, M.; La Rosa, F.; Piancone, F.; Zoppis, M.; Marventano, I.; Calabrese, E.; Rainone, V.; Nemni, R.; Mancuso, R.; Clerici, M. The NLRP3 and NLRP1 inflammasomes are activated in Alzheimer’s disease. Mol. Neurodegener. 2016, 11, 23. [Google Scholar] [CrossRef] [PubMed]
  90. Heneka, M.T.; Kummer, M.P.; Stutz, A.; Delekate, A.; Schwartz, S.; Vieira-Saecker, A.; Griep, A.; Axt, D.; Remus, A.; Tzeng, T.C.; et al. NLRP3 is activated in Alzheimer’s disease and contributes to pathology in APP/PS1 mice. Nature 2013, 493, 674–678. [Google Scholar] [CrossRef] [PubMed]
  91. Bradshaw, E.M.; Chibnik, L.B.; Keenan, B.T.; Ottoboni, L.; Raj, T.; Tang, A.; Rosenkrantz, L.L.; Imboywa, S.; Lee, M.; Von Korff, A.; et al. CD33 Alzheimer’s disease locus: Altered monocyte function and amyloid biology. Nat. Neurosci. 2013, 16, 848–850. [Google Scholar] [CrossRef] [PubMed]
  92. Liu, G.; Jiang, Q. Alzheimer’s disease CD33 rs3865444 variant does not contribute to cognitive performance. Proc. Natl. Acad. Sci. USA 2016, 113, E1589–E1590. [Google Scholar] [CrossRef] [PubMed]
  93. Griciuc, A.; Serrano-Pozo, A.; Parrado, A.R.; Lesinski, A.N.; Asselin, C.N.; Mullin, K.; Hooli, B.; Choi, S.H.; Hyman, B.T.; Tanzi, R.E. Alzheimer’s disease risk gene CD33 inhibits microglial uptake of amyloid beta. Neuron 2013, 78, 631–643. [Google Scholar] [CrossRef] [PubMed]
  94. Hickman, S.E.; Allison, E.K.; El Khoury, J. Microglial dysfunction and defective beta-amyloid clearance pathways in aging Alzheimer’s disease mice. J. Neurosci. 2008, 28, 8354–8360. [Google Scholar] [CrossRef] [PubMed]
  95. Mawuenyega, K.G.; Sigurdson, W.; Ovod, V.; Munsell, L.; Kasten, T.; Morris, J.C.; Yarasheski, K.E.; Bateman, R.J. Decreased clearance of CNS beta-amyloid in Alzheimer’s disease. Science 2010, 330, 1774. [Google Scholar] [CrossRef]
  96. Xue, J.; Schmidt, S.V.; Sander, J.; Draffehn, A.; Krebs, W.; Quester, I.; De Nardo, D.; Gohel, T.D.; Emde, M.; Schmidleithner, L.; et al. Transcriptome-based network analysis reveals a spectrum model of human macrophage activation. Immunity 2014, 40, 274–288. [Google Scholar] [CrossRef]
  97. Ransohoff, R.M. A polarizing question: Do M1 and M2 microglia exist? Nat. Neurosci. 2016, 19, 987–991. [Google Scholar] [CrossRef]
  98. Orihuela, R.; McPherson, C.A.; Harry, G.J. Microglial M1/M2 polarization and metabolic states. Br. J. Pharmacol. 2016, 173, 649–665. [Google Scholar] [CrossRef]
  99. Weldon, D.T.; Rogers, S.D.; Ghilardi, J.R.; Finke, M.P.; Cleary, J.P.; O’Hare, E.; Esler, W.P.; Maggio, J.E.; Mantyh, P.W. Fibrillar beta-amyloid induces microglial phagocytosis, expression of inducible nitric oxide synthase, and loss of a select population of neurons in the rat CNS in vivo. J. Neurosci. 1998, 18, 2161–2173. [Google Scholar] [CrossRef]
  100. Koenigsknecht-Talboo, J.; Landreth, G.E. Microglial phagocytosis induced by fibrillar beta-amyloid and IgGs are differentially regulated by proinflammatory cytokines. J. Neurosci. 2005, 25, 8240–8249. [Google Scholar] [CrossRef]
  101. Yamamoto, M.; Kiyota, T.; Walsh, S.M.; Liu, J.; Kipnis, J.; Ikezu, T. Cytokine-mediated inhibition of fibrillar amyloid-beta peptide degradation by human mononuclear phagocytes. J. Immunol. 2008, 181, 3877–3886. [Google Scholar] [CrossRef] [PubMed]
  102. Michelucci, A.; Heurtaux, T.; Grandbarbe, L.; Morga, E.; Heuschling, P. Characterization of the microglial phenotype under specific pro-inflammatory and anti-inflammatory conditions: Effects of oligomeric and fibrillar amyloid-beta. J. Neuroimmunol. 2009, 210, 3–12. [Google Scholar] [CrossRef] [PubMed]
  103. Dempsey, C.; Rubio Araiz, A.; Bryson, K.J.; Finucane, O.; Larkin, C.; Mills, E.L.; Robertson, A.A.; Cooper, M.A.; O’Neill, L.A.; Lynch, M.A. Inhibiting the NLRP3 inflammasome with MCC950 promotes non-phlogistic clearance of amyloid-beta and cognitive function in APP/PS1 mice. Brain Behav. Immun. 2017, 61, 306–316. [Google Scholar] [CrossRef] [PubMed]
  104. Flores, J.; Noel, A.; Foveau, B.; Lynham, J.; Lecrux, C.; LeBlanc, A.C. Caspase-1 inhibition alleviates cognitive impairment and neuropathology in an Alzheimer’s disease mouse model. Nat. Commun. 2018, 9, 3916. [Google Scholar] [CrossRef] [PubMed]
  105. Tan, M.S.; Yu, J.T.; Jiang, T.; Zhu, X.C.; Wang, H.F.; Zhang, W.; Wang, Y.L.; Jiang, W.; Tan, L. NLRP3 polymorphisms are associated with late-onset Alzheimer’s disease in Han Chinese. J. Neuroimmunol. 2013, 265, 91–95. [Google Scholar] [CrossRef] [PubMed]
  106. Tan, M.S.; Yu, J.T.; Jiang, T.; Zhu, X.C.; Tan, L. The NLRP3 inflammasome in Alzheimer’s disease. Mol. Neurobiol. 2013, 48, 875–882. [Google Scholar] [CrossRef]
  107. Crismon, M.L. Tacrine: First drug approved for Alzheimer’s disease. Ann. Pharmacother. 1994, 28, 744–751. [Google Scholar] [CrossRef]
  108. Pan, S.Y.; Guo, B.F.; Zhang, Y.; Yu, Q.; Yu, Z.L.; Dong, H.; Ye, Y.; Han, Y.F.; Ko, K.M. Tacrine treatment at high dose suppresses the recognition memory in juvenile and adult mice with attention to hepatotoxicity. Basic Clin. Pharmacol. Toxicol. 2011, 108, 421–427. [Google Scholar] [CrossRef]
  109. Galisteo, M.; Rissel, M.; Sergent, O.; Chevanne, M.; Cillard, J.; Guillouzo, A.; Lagadic-Gossmann, D. Hepatotoxicity of tacrine: Occurrence of membrane fluidity alterations without involvement of lipid peroxidation. J. Pharmacol. Exp. Ther. 2000, 294, 160–167. [Google Scholar]
  110. Gutzmann, H.; Kuhl, K.P.; Hadler, D.; Rapp, M.A. Safety and efficacy of idebenone versus tacrine in patients with Alzheimer’s disease: Results of a randomized, double-blind, parallel-group multicenter study. Pharmacopsychiatry 2002, 35, 12–18. [Google Scholar] [CrossRef]
  111. Gracon, S.I.; Knapp, M.J.; Berghoff, W.G.; Pierce, M.; DeJong, R.; Lobbestael, S.J.; Symons, J.; Dombey, S.L.; Luscombe, F.A.; Kraemer, D. Safety of tacrine: Clinical trials, treatment IND, and postmarketing experience. Alzheimer Dis. Assoc. Disord. 1998, 12, 93–101. [Google Scholar] [CrossRef] [PubMed]
  112. Grossberg, G.T. Effect of rivastigmine in the treatment of behavioral disturbances associated with dementia: Review of neuropsychiatric impairment in Alzheimer’s disease. Curr. Med. Res. Opin. 2005, 21, 1631–1639. [Google Scholar] [CrossRef] [PubMed]
  113. Rosler, M.; Anand, R.; Cicin-Sain, A.; Gauthier, S.; Agid, Y.; Dal-Bianco, P.; Stahelin, H.B.; Hartman, R.; Gharabawi, M. Efficacy and safety of rivastigmine in patients with Alzheimer’s disease: International randomised controlled trial. BMJ 1999, 318, 633–638. [Google Scholar] [CrossRef] [PubMed]
  114. Olin, J.; Schneider, L. Galantamine for Alzheimer’s disease. Cochrane Database Syst. Rev. 2001, 3, CD001747. [Google Scholar] [CrossRef]
  115. Pirttila, T.; Wilcock, G.; Truyen, L.; Damaraju, C.V. Long-term efficacy and safety of galantamine in patients with mild-to-moderate Alzheimer’s disease: Multicenter trial. Eur. J. Neurol. 2004, 11, 734–741. [Google Scholar] [CrossRef] [PubMed]
  116. Benjamin, B.; Burns, A. Donepezil for Alzheimer’s disease. Expert Rev. Neurother. 2007, 7, 1243–1249. [Google Scholar] [CrossRef] [PubMed]
  117. Winblad, B.; Kilander, L.; Eriksson, S.; Minthon, L.; Batsman, S.; Wetterholm, A.L.; Jansson-Blixt, C.; Haglund, A.; Severe Alzheimer’s Disease Study Group. Donepezil in patients with severe Alzheimer’s disease: Double-blind, parallel-group, placebo-controlled study. Lancet 2006, 367, 1057–1065. [Google Scholar] [CrossRef]
  118. Birks, J. Cholinesterase inhibitors for Alzheimer’s disease. Cochrane Database Syst. Rev. 2006, 1, CD005593. [Google Scholar] [CrossRef]
  119. Ali, T.B.; Schleret, T.R.; Reilly, B.M.; Chen, W.Y.; Abagyan, R. Adverse Effects of Cholinesterase Inhibitors in Dementia, According to the Pharmacovigilance Databases of the United-States and Canada. PLoS ONE 2015, 10, e0144337. [Google Scholar] [CrossRef] [PubMed]
  120. Inglis, F. The tolerability and safety of cholinesterase inhibitors in the treatment of dementia. Int. J. Clin. Pract. Suppl. 2002, 45–63. [Google Scholar]
  121. Bleich, S.; Wiltfang, J.; Kornhuber, J. Memantine in moderate-to-severe Alzheimer’s disease. New Engl. J. Med. 2003, 349, 609–610. [Google Scholar] [CrossRef] [PubMed]
  122. Schneider, L.S.; Dagerman, K.S.; Higgins, J.P.; McShane, R. Lack of evidence for the efficacy of memantine in mild Alzheimer disease. Arch. Neurol. 2011, 68, 991–998. [Google Scholar] [CrossRef] [PubMed]
  123. Tan, M.G.; Shear, N.H.; Walsh, S.; Alhusayen, R. Acitretin. J. Cutan. Med. Surg. 2017, 21, 48–53. [Google Scholar] [CrossRef] [PubMed]
  124. Guenther, L.C.; Kunynetz, R.; Lynde, C.W.; Sibbald, R.G.; Toole, J.; Vender, R.; Zip, C. Acitretin Use in Dermatology. J. Cutan. Med. Surg. 2017, 21, 2S–S12. [Google Scholar] [CrossRef] [PubMed]
  125. Heath, M.S.; Sahni, D.R.; Curry, Z.A.; Feldman, S.R. Pharmacokinetics of tazarotene and acitretin in psoriasis. Expert Opin. Drug Metabol. Toxicol. 2018, 14, 919–927. [Google Scholar] [CrossRef] [PubMed]
  126. Endres, K.; Fahrenholz, F.; Lotz, J.; Hiemke, C.; Teipel, S.; Lieb, K.; Tuscher, O.; Fellgiebel, A. Increased CSF APPs-alpha levels in patients with Alzheimer disease treated with acitretin. Neurology 2014, 83, 1930–1935. [Google Scholar] [CrossRef] [PubMed]
  127. Xicota, L.; Rodriguez-Morato, J.; Dierssen, M.; de la Torre, R. Potential Role of (−)-Epigallocatechin-3-Gallate (EGCG) in the Secondary Prevention of Alzheimer Disease. Curr. Drug Targets 2017, 18, 174–195. [Google Scholar] [CrossRef] [PubMed]
  128. Singh, N.A.; Mandal, A.K.; Khan, Z.A. Potential neuroprotective properties of epigallocatechin-3-gallate (EGCG). Nutr. J. 2016, 15, 60. [Google Scholar] [CrossRef]
  129. Siopi, E.; Llufriu-Daben, G.; Cho, A.H.; Vidal-Lletjos, S.; Plotkine, M.; Marchand-Leroux, C.; Jafarian-Tehrani, M. Etazolate, an alpha-secretase activator, reduces neuroinflammation and offers persistent neuroprotection following traumatic brain injury in mice. Neuropharmacology 2013, 67, 183–192. [Google Scholar] [CrossRef] [PubMed]
  130. Marcade, M.; Bourdin, J.; Loiseau, N.; Peillon, H.; Rayer, A.; Drouin, D.; Schweighoffer, F.; Desire, L. Etazolate, a neuroprotective drug linking GABA(A) receptor pharmacology to amyloid precursor protein processing. J. Neurochem. 2008, 106, 392–404. [Google Scholar] [CrossRef] [PubMed]
  131. Cebers, G.; Alexander, R.C.; Haeberlein, S.B.; Han, D.; Goldwater, R.; Ereshefsky, L.; Olsson, T.; Ye, N.; Rosen, L.; Russell, M.; et al. AZD3293: Pharmacokinetic and Pharmacodynamic Effects in Healthy Subjects and Patients with Alzheimer’s Disease. J. Alzheimer’s Dis. JAD 2017, 55, 1039–1053. [Google Scholar] [CrossRef] [PubMed]
  132. Eketjall, S.; Janson, J.; Kaspersson, K.; Bogstedt, A.; Jeppsson, F.; Falting, J.; Haeberlein, S.B.; Kugler, A.R.; Alexander, R.C.; Cebers, G. AZD3293: A Novel, Orally Active BACE1 Inhibitor with High Potency and Permeability and Markedly Slow Off-Rate Kinetics. J. Alzheimer’s Dis. JAD 2016, 50, 1109–1123. [Google Scholar] [CrossRef] [PubMed]
  133. Burki, T. Alzheimer’s disease research: The future of BACE inhibitors. Lancet 2018, 391, 2486. [Google Scholar] [CrossRef]
  134. Blume, T.; Filser, S.; Jaworska, A.; Blain, J.F.; Koenig, G.; Moschke, K.; Lichtenthaler, S.F.; Herms, J. BACE1 Inhibitor MK-8931 Alters Formation but Not Stability of Dendritic Spines. Front. Aging Neurosci. 2018, 10, 229. [Google Scholar] [CrossRef] [PubMed]
  135. Thaisrivongs, D.A.; Morris, W.J.; Tan, L.; Song, Z.J.; Lyons, T.W.; Waldman, J.H.; Naber, J.R.; Chen, W.; Chen, L.; Zhang, B.; et al. A Next Generation Synthesis of BACE1 Inhibitor Verubecestat (MK-8931). Org. Lett. 2018, 20, 1568–1571. [Google Scholar] [CrossRef]
  136. Scott, J.D.; Li, S.W.; Brunskill, A.P.; Chen, X.; Cox, K.; Cumming, J.N.; Forman, M.; Gilbert, E.J.; Hodgson, R.A.; Hyde, L.A.; et al. Discovery of the 3-Imino-1,2,4-thiadiazinane 1,1-Dioxide Derivative Verubecestat (MK-8931)-A beta-Site Amyloid Precursor Protein Cleaving Enzyme 1 Inhibitor for the Treatment of Alzheimer’s Disease. J. Med. Chem. 2016, 59, 10435–10450. [Google Scholar] [CrossRef]
  137. Kennedy, M.E.; Stamford, A.W.; Chen, X.; Cox, K.; Cumming, J.N.; Dockendorf, M.F.; Egan, M.; Ereshefsky, L.; Hodgson, R.A.; Hyde, L.A.; et al. The BACE1 inhibitor verubecestat (MK-8931) reduces CNS beta-amyloid in animal models and in Alzheimer’s disease patients. Sci. Transl. Med. 2016, 8, 363ra150. [Google Scholar] [CrossRef]
  138. Timmers, M.; Streffer, J.R.; Russu, A.; Tominaga, Y.; Shimizu, H.; Shiraishi, A.; Tatikola, K.; Smekens, P.; Borjesson-Hanson, A.; Andreasen, N.; et al. Pharmacodynamics of atabecestat (JNJ-54861911), an oral BACE1 inhibitor in patients with early Alzheimer’s disease: Randomized, double-blind, placebo-controlled study. Alzheimer’s Res. Ther. 2018, 10, 85. [Google Scholar] [CrossRef]
  139. Timmers, M.; Sinha, V.; Darpo, B.; Smith, B.; Brown, R.; Xue, H.; Ferber, G.; Streffer, J.; Russu, A.; Tritsmans, L.; et al. Evaluating Potential QT Effects of JNJ-54861911, a BACE Inhibitor in Single- and Multiple-Ascending Dose Studies, and a Thorough QT Trial With Additional Retrospective Confirmation, Using Concentration-QTc Analysis. J. Clin. Pharmacol. 2018, 58, 952–964. [Google Scholar] [CrossRef]
  140. Timmers, M.; Van Broeck, B.; Ramael, S.; Slemmon, J.; De Waepenaert, K.; Russu, A.; Bogert, J.; Stieltjes, H.; Shaw, L.M.; Engelborghs, S.; et al. Profiling the dynamics of CSF and plasma Abeta reduction after treatment with JNJ-54861911, a potent oral BACE inhibitor. Alzheimers Dement. (N. Y.) 2016, 2, 202–212. [Google Scholar] [CrossRef]
  141. Neumann, U.; Ufer, M.; Jacobson, L.H.; Rouzade-Dominguez, M.L.; Huledal, G.; Kolly, C.; Luond, R.M.; Machauer, R.; Veenstra, S.J.; Hurth, K.; et al. The BACE-1 inhibitor CNP520 for prevention trials in Alzheimer’s disease. EMBO Mol. Med. 2018, 10, e9316. [Google Scholar] [CrossRef] [PubMed]
  142. Lopez Lopez, C.; Caputo, A.; Liu, F.; Riviere, M.E.; Rouzade-Dominguez, M.L.; Thomas, R.G.; Langbaum, J.B.; Lenz, R.; Reiman, E.M.; Graf, A.; et al. The Alzheimer’s Prevention Initiative Generation Program: Evaluating CNP520 Efficacy in the Prevention of Alzheimer’s Disease. J. Prev. Alzheimer’s Dis. 2017, 4, 242–246. [Google Scholar] [CrossRef]
  143. Folch, J.; Petrov, D.; Ettcheto, M.; Abad, S.; Sanchez-Lopez, E.; Garcia, M.L.; Olloquequi, J.; Beas-Zarate, C.; Auladell, C.; Camins, A. Current Research Therapeutic Strategies for Alzheimer’s Disease Treatment. Neural Plast. 2016, 2016, 8501693. [Google Scholar] [CrossRef] [PubMed]
  144. Dockens, R.; Wang, J.S.; Castaneda, L.; Sverdlov, O.; Huang, S.P.; Slemmon, R.; Gu, H.; Wong, O.; Li, H.; Berman, R.M.; et al. A placebo-controlled, multiple ascending dose study to evaluate the safety, pharmacokinetics and pharmacodynamics of avagacestat (BMS-708163) in healthy young and elderly subjects. Clin. Pharmacokinet. 2012, 51, 681–693. [Google Scholar] [CrossRef] [PubMed]
  145. Tong, G.; Castaneda, L.; Wang, J.S.; Sverdlov, O.; Huang, S.P.; Slemmon, R.; Gu, H.; Wong, O.; Li, H.; Berman, R.M.; et al. Effects of single doses of avagacestat (BMS-708163) on cerebrospinal fluid Abeta levels in healthy young men. Clin. Drug Investig. 2012, 32, 761–769. [Google Scholar] [CrossRef] [PubMed]
  146. Coric, V.; van Dyck, C.H.; Salloway, S.; Andreasen, N.; Brody, M.; Richter, R.W.; Soininen, H.; Thein, S.; Shiovitz, T.; Pilcher, G.; et al. Safety and tolerability of the gamma-secretase inhibitor avagacestat in a phase 2 study of mild to moderate Alzheimer disease. Arch. Neurol. 2012, 69, 1430–1440. [Google Scholar] [CrossRef]
  147. Matlack, K.E.; Tardiff, D.F.; Narayan, P.; Hamamichi, S.; Caldwell, K.A.; Caldwell, G.A.; Lindquist, S. Clioquinol promotes the degradation of metal-dependent amyloid-beta (Abeta) oligomers to restore endocytosis and ameliorate Abeta toxicity. Proc. Natl. Acad. Sci. USA 2014, 111, 4013–4018. [Google Scholar] [CrossRef] [PubMed]
  148. Liang, E.; Garzone, P.; Cedarbaum, J.M.; Koller, M.; Tran, T.; Xu, V.; Ross, B.; Jhee, S.S.; Ereshefsky, L.; Pastrak, A.; et al. Pharmacokinetic Profile of Orally Administered Scyllo-Inositol (Elnd005) in Plasma, Cerebrospinal Fluid and Brain, and Corresponding Effect on Amyloid-Beta in Healthy Subjects. Clin. Pharmacol. Drug Dev. 2013, 2, 186–194. [Google Scholar] [CrossRef] [PubMed]
  149. Salloway, S.; Sperling, R.; Keren, R.; Porsteinsson, A.P.; van Dyck, C.H.; Tariot, P.N.; Gilman, S.; Arnold, D.; Abushakra, S.; Hernandez, C.; et al. A phase 2 randomized trial of ELND005, scyllo-inositol, in mild to moderate Alzheimer disease. Neurology 2011, 77, 1253–1262. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Caltagirone, C.; Ferrannini, L.; Marchionni, N.; Nappi, G.; Scapagnini, G.; Trabucchi, M. The potential protective effect of tramiprosate (homotaurine) against Alzheimer’s disease: A review. Aging Clin. Exp. Res. 2012, 24, 580–587. [Google Scholar] [CrossRef] [PubMed]
  151. Aisen, P.S.; Gauthier, S.; Ferris, S.H.; Saumier, D.; Haine, D.; Garceau, D.; Duong, A.; Suhy, J.; Oh, J.; Lau, W.C.; et al. Tramiprosate in mild-to-moderate Alzheimer’s disease—A randomized, double-blind, placebo-controlled, multi-centre study (the Alphase Study). Arch. Med. Sci. AMS 2011, 7, 102–111. [Google Scholar] [CrossRef] [PubMed]
  152. Gervais, F.; Paquette, J.; Morissette, C.; Krzywkowski, P.; Yu, M.; Azzi, M.; Lacombe, D.; Kong, X.; Aman, A.; Laurin, J.; et al. Targeting soluble Abeta peptide with Tramiprosate for the treatment of brain amyloidosis. Neurobiol. Aging 2007, 28, 537–547. [Google Scholar] [CrossRef] [PubMed]
  153. Gauthier, S.; Aisen, P.S.; Ferris, S.H.; Saumier, D.; Duong, A.; Haine, D.; Garceau, D.; Suhy, J.; Oh, J.; Lau, W.; et al. Effect of tramiprosate in patients with mild-to-moderate Alzheimer’s disease: Exploratory analyses of the MRI sub-group of the Alphase study. J. Nutr. Health Aging 2009, 13, 550–557. [Google Scholar] [CrossRef]
  154. Sabbagh, M.N. Clinical Effects of Oral Tramiprosate in APOE4/4 Homozygous Patients with Mild Alzheimer’s Disease Suggest Disease Modification. J. Prev. Alzheimer’s Dis. 2017, 4, 136–137. [Google Scholar] [CrossRef]
  155. Kocis, P.; Tolar, M.; Yu, J.; Sinko, W.; Ray, S.; Blennow, K.; Fillit, H.; Hey, J.A. Elucidating the Abeta42 Anti-Aggregation Mechanism of Action of Tramiprosate in Alzheimer’s Disease: Integrating Molecular Analytical Methods, Pharmacokinetic and Clinical Data. CNS Drugs 2017, 31, 495–509. [Google Scholar] [CrossRef] [PubMed]
  156. Abushakra, S.; Porsteinsson, A.; Scheltens, P.; Sadowsky, C.; Vellas, B.; Cummings, J.; Gauthier, S.; Hey, J.A.; Power, A.; Wang, P.; et al. Clinical Effects of Tramiprosate in APOE4/4 Homozygous Patients with Mild Alzheimer’s Disease Suggest Disease Modification Potential. J. Prev. Alzheimer’s Dis. 2017, 4, 149–156. [Google Scholar] [CrossRef]
  157. Abushakra, S.; Porsteinsson, A.; Vellas, B.; Cummings, J.; Gauthier, S.; Hey, J.A.; Power, A.; Hendrix, S.; Wang, P.; Shen, L.; et al. Clinical Benefits of Tramiprosate in Alzheimer’s Disease Are Associated with Higher Number of APOE4 Alleles: The “APOE4 Gene-Dose Effect”. J. Prev. Alzheimer’s Dis. 2016, 3, 219–228. [Google Scholar] [CrossRef]
  158. Liu, Y.; Xu, L.P.; Wang, Q.; Yang, B.; Zhang, X. Synergistic Inhibitory Effect of GQDs-Tramiprosate Covalent Binding on Amyloid Aggregation. ACS Chem. Neurosci. 2018, 9, 817–823. [Google Scholar] [CrossRef]
  159. Hey, J.A.; Yu, J.Y.; Versavel, M.; Abushakra, S.; Kocis, P.; Power, A.; Kaplan, P.L.; Amedio, J.; Tolar, M. Clinical Pharmacokinetics and Safety of ALZ-801, a Novel Prodrug of Tramiprosate in Development for the Treatment of Alzheimer’s Disease. Clin. Pharmacokinet. 2018, 57, 315–333. [Google Scholar] [CrossRef]
  160. Hey, J.A.; Kocis, P.; Hort, J.; Abushakra, S.; Power, A.; Vyhnalek, M.; Yu, J.Y.; Tolar, M. Discovery and Identification of an Endogenous Metabolite of Tramiprosate and Its Prodrug ALZ-801 that Inhibits Beta Amyloid Oligomer Formation in the Human Brain. CNS Drugs 2018, 32, 849–861. [Google Scholar] [CrossRef]
  161. Cummings, J.; Lee, G.; Ritter, A.; Zhong, K. Alzheimer’s disease drug development pipeline: 2018. Alzheimers Dement. (N. Y.) 2018, 4, 195–214. [Google Scholar] [CrossRef] [PubMed]
  162. 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] [PubMed]
  163. Holmes, C.; Boche, D.; Wilkinson, D.; Yadegarfar, G.; Hopkins, V.; Bayer, A.; Jones, R.W.; Bullock, R.; Love, S.; Neal, J.W.; et al. Long-term effects of Abeta42 immunisation in Alzheimer’s disease: Follow-up of a randomised, placebo-controlled phase I trial. Lancet 2008, 372, 216–223. [Google Scholar] [CrossRef]
  164. Hull, M.; Sadowsky, C.; Arai, H.; Le Prince Leterme, G.; Holstein, A.; Booth, K.; Peng, Y.; Yoshiyama, T.; Suzuki, H.; Ketter, N.; et al. Long-Term Extensions of Randomized Vaccination Trials of ACC-001 and QS-21 in Mild to Moderate Alzheimer’s Disease. Curr. Alzheimer Res. 2017, 14, 696–708. [Google Scholar] [CrossRef] [PubMed]
  165. Pasquier, F.; Sadowsky, C.; Holstein, A.; Leterme Gle, P.; Peng, Y.; Jackson, N.; Fox, N.C.; Ketter, N.; Liu, E.; Ryan, J.M.; et al. Two Phase 2 Multiple Ascending-Dose Studies of Vanutide Cridificar (ACC-001) and QS-21 Adjuvant in Mild-to-Moderate Alzheimer’s Disease. J. Alzheimer’s Dis. JAD 2016, 51, 1131–1143. [Google Scholar] [CrossRef]
  166. van Dyck, C.H.; Sadowsky, C.; Le Prince Leterme, G.; Booth, K.; Peng, Y.; Marek, K.; Ketter, N.; Liu, E.; Wyman, B.T.; Jackson, N.; et al. Vanutide Cridificar (ACC-001) and QS-21 Adjuvant in Individuals with Early Alzheimer’s Disease: Amyloid Imaging Positron Emission Tomography and Safety Results from a Phase 2 Study. J. Prev. Alzheimer’s Dis. 2016, 3, 75–84. [Google Scholar] [CrossRef]
  167. Tayeb, H.O.; Murray, E.D.; Price, B.H.; Tarazi, F.I. Bapineuzumab and solanezumab for Alzheimer’s disease: Is the ‘amyloid cascade hypothesis’ still alive? Expert Opin. Biol. Ther. 2013, 13, 1075–1084. [Google Scholar] [CrossRef]
  168. Doody, R.S.; Farlow, M.; Aisen, P.S.; Alzheimer’s Disease Cooperative Study Data, A.; Publication, C. Phase 3 trials of solanezumab and bapineuzumab for Alzheimer’s disease. N. Engl. J. Med. 2014, 370, 1460. [Google Scholar] [CrossRef]
  169. La Porte, S.L.; Bollini, S.S.; Lanz, T.A.; Abdiche, Y.N.; Rusnak, A.S.; Ho, W.H.; Kobayashi, D.; Harrabi, O.; Pappas, D.; Mina, E.W.; et al. Structural basis of C-terminal beta-amyloid peptide binding by the antibody ponezumab for the treatment of Alzheimer’s disease. J. Mol. Biol. 2012, 421, 525–536. [Google Scholar] [CrossRef]
  170. Landen, J.W.; Zhao, Q.; Cohen, S.; Borrie, M.; Woodward, M.; Billing, C.B., Jr.; Bales, K.; Alvey, C.; McCush, F.; Yang, J.; et al. Safety and pharmacology of a single intravenous dose of ponezumab in subjects with mild-to-moderate Alzheimer disease: A phase I, randomized, placebo-controlled, double-blind, dose-escalation study. Clin. Neuropharmacol. 2013, 36, 14–23. [Google Scholar] [CrossRef]
  171. Landen, J.W.; Cohen, S.; Billing, C.B., Jr.; Cronenberger, C.; Styren, S.; Burstein, A.H.; Sattler, C.; Lee, J.H.; Jack, C.R., Jr.; Kantarci, K.; et al. Multiple-dose ponezumab for mild-to-moderate Alzheimer’s disease: Safety and efficacy. Alzheimers Dement. (N. Y.) 2017, 3, 339–347. [Google Scholar] [CrossRef] [PubMed]
  172. Landen, J.W.; Andreasen, N.; Cronenberger, C.L.; Schwartz, P.F.; Borjesson-Hanson, A.; Ostlund, H.; Sattler, C.A.; Binneman, B.; Bednar, M.M. Ponezumab in mild-to-moderate Alzheimer’s disease: Randomized phase II PET-PIB study. Alzheimers Dement. (N. Y.) 2017, 3, 393–401. [Google Scholar] [CrossRef] [PubMed]
  173. Andreasen, N.; Simeoni, M.; Ostlund, H.; Lisjo, P.I.; Fladby, T.; Loercher, A.E.; Byrne, G.J.; Murray, F.; Scott-Stevens, P.T.; Wallin, A.; et al. First administration of the Fc-attenuated anti-beta amyloid antibody GSK933776 to patients with mild Alzheimer’s disease: A randomized, placebo-controlled study. PLoS ONE 2015, 10, e0098153. [Google Scholar] [CrossRef] [PubMed]
  174. Leyhe, T.; Andreasen, N.; Simeoni, M.; Reich, A.; von Arnim, C.A.; Tong, X.; Yeo, A.; Khan, S.; Loercher, A.; Chalker, M.; et al. Modulation of beta-amyloid by a single dose of GSK933776 in patients with mild Alzheimer’s disease: A phase I study. Alzheimer’s Res. Ther. 2014, 6, 19. [Google Scholar] [CrossRef] [PubMed]
  175. Ferrero, J.; Williams, L.; Stella, H.; Leitermann, K.; Mikulskis, A.; O’Gorman, J.; Sevigny, J. First-in-human, double-blind, placebo-controlled, single-dose escalation study of aducanumab (BIIB037) in mild-to-moderate Alzheimer’s disease. Alzheimers Dement. (N. Y.) 2016, 2, 169–176. [Google Scholar] [CrossRef] [PubMed]
  176. Arndt, J.W.; Qian, F.; Smith, B.A.; Quan, C.; Kilambi, K.P.; Bush, M.W.; Walz, T.; Pepinsky, R.B.; Bussiere, T.; Hamann, S.; et al. Structural and kinetic basis for the selectivity of aducanumab for aggregated forms of amyloid-beta. Sci. Rep. 2018, 8, 6412. [Google Scholar] [CrossRef]
  177. Gamage, K.K.; Kumar, S. Aducanumab Therapy Ameliorates Calcium Overload in a Mouse Model of Alzheimer’s Disease. J. Neurosci. 2017, 37, 4430–4432. [Google Scholar] [CrossRef]
  178. Kastanenka, K.V.; Bussiere, T.; Shakerdge, N.; Qian, F.; Weinreb, P.H.; Rhodes, K.; Bacskai, B.J. Immunotherapy with Aducanumab Restores Calcium Homeostasis in Tg2576 Mice. J. Neurosci. 2016, 36, 12549–12558. [Google Scholar] [CrossRef]
  179. Budd Haeberlein, S.; O’Gorman, J.; Chiao, P.; Bussiere, T.; von Rosenstiel, P.; Tian, Y.; Zhu, Y.; von Hehn, C.; Gheuens, S.; Skordos, L.; et al. Clinical Development of Aducanumab, an Anti-Abeta Human Monoclonal Antibody Being Investigated for the Treatment of Early Alzheimer’s Disease. J. Prev. Alzheimer’s Dis. 2017, 4, 255–263. [Google Scholar] [CrossRef]
  180. Sevigny, J.; Chiao, P.; Bussiere, T.; Weinreb, P.H.; Williams, L.; Maier, M.; Dunstan, R.; Salloway, S.; Chen, T.; Ling, Y.; et al. The antibody aducanumab reduces Abeta plaques in Alzheimer’s disease. Nature 2016, 537, 50–56. [Google Scholar] [CrossRef]
  181. Chiao, P.; Bedell, B.J.; Avants, B.; Zijdenbos, A.P.; Grand’Maison, M.; O’Neill, P.; O’Gorman, J.; Chen, T.; Koeppe, R. Impact of Reference/Target Region Selection on Amyloid PET Standard Uptake Value Ratios in the Phase 1b PRIME Study of Aducanumab. J. Nucl. Med. 2018. [Google Scholar] [CrossRef]
  182. Ultsch, M.; Li, B.; Maurer, T.; Mathieu, M.; Adolfsson, O.; Muhs, A.; Pfeifer, A.; Pihlgren, M.; Bainbridge, T.W.; Reichelt, M.; et al. Structure of Crenezumab Complex with Abeta Shows Loss of beta-Hairpin. Sci. Rep. 2016, 6, 39374. [Google Scholar] [CrossRef]
  183. Salloway, S.; Honigberg, L.A.; Cho, W.; Ward, M.; Friesenhahn, M.; Brunstein, F.; Quartino, A.; Clayton, D.; Mortensen, D.; Bittner, T.; et al. Amyloid positron emission tomography and cerebrospinal fluid results from a crenezumab anti-amyloid-beta antibody double-blind, placebo-controlled, randomized phase II study in mild-to-moderate Alzheimer’s disease (BLAZE). Alzheimer’s Res. Ther. 2018, 10, 96. [Google Scholar] [CrossRef]
  184. Cummings, J.L.; Cohen, S.; van Dyck, C.H.; Brody, M.; Curtis, C.; Cho, W.; Ward, M.; Friesenhahn, M.; Rabe, C.; Brunstein, F.; et al. ABBY: A phase 2 randomized trial of crenezumab in mild to moderate Alzheimer disease. Neurology 2018, 90, e1889–e1897. [Google Scholar] [CrossRef] [PubMed]
  185. Bohrmann, B.; Baumann, K.; Benz, J.; Gerber, F.; Huber, W.; Knoflach, F.; Messer, J.; Oroszlan, K.; Rauchenberger, R.; Richter, W.F.; et al. Gantenerumab: A novel human anti-Abeta antibody demonstrates sustained cerebral amyloid-beta binding and elicits cell-mediated removal of human amyloid-beta. J. Alzheimer’s Dis. JAD 2012, 28, 49–69. [Google Scholar] [CrossRef]
  186. Ostrowitzki, S.; Deptula, D.; Thurfjell, L.; Barkhof, F.; Bohrmann, B.; Brooks, D.J.; Klunk, W.E.; Ashford, E.; Yoo, K.; Xu, Z.X.; et al. Mechanism of amyloid removal in patients with Alzheimer disease treated with gantenerumab. Arch. Neurol. 2012, 69, 198–207. [Google Scholar] [CrossRef]
  187. Delrieu, J.; Ousset, P.J.; Vellas, B. Gantenerumab for the treatment of Alzheimer’s disease. Expert Opin. Biol. Ther. 2012, 12, 1077–1086. [Google Scholar] [CrossRef]
  188. Novakovic, D.; Feligioni, M.; Scaccianoce, S.; Caruso, A.; Piccinin, S.; Schepisi, C.; Errico, F.; Mercuri, N.B.; Nicoletti, F.; Nistico, R. Profile of gantenerumab and its potential in the treatment of Alzheimer’s disease. Drug Des. Dev. Ther. 2013, 7, 1359–1364. [Google Scholar] [CrossRef]
  189. 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. Alzheimer’s Res. Ther. 2017, 9, 95. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Baddeley, T.C.; McCaffrey, J.; Storey, J.M.; Cheung, J.K.; Melis, V.; Horsley, D.; Harrington, C.R.; Wischik, C.M. Complex disposition of methylthioninium redox forms determines efficacy in tau aggregation inhibitor therapy for Alzheimer’s disease. J. Pharmacol. Exp. Ther. 2015, 352, 110–118. [Google Scholar] [CrossRef]
  191. Gauthier, S.; Feldman, H.H.; Schneider, L.S.; Wilcock, G.K.; Frisoni, G.B.; Hardlund, J.H.; Moebius, H.J.; Bentham, P.; Kook, K.A.; Wischik, D.J.; et al. Efficacy and safety of tau-aggregation inhibitor therapy in patients with mild or moderate Alzheimer’s disease: A randomised, controlled, double-blind, parallel-arm, phase 3 trial. Lancet 2016, 388, 2873–2884. [Google Scholar] [CrossRef]
  192. Wilcock, G.K.; Gauthier, S.; Frisoni, G.B.; Jia, J.; Hardlund, J.H.; Moebius, H.J.; Bentham, P.; Kook, K.A.; Schelter, B.O.; Wischik, D.J.; et al. Potential of Low Dose Leuco-Methylthioninium Bis(Hydromethanesulphonate) (LMTM) Monotherapy for Treatment of Mild Alzheimer’s Disease: Cohort Analysis as Modified Primary Outcome in a Phase III Clinical Trial. J. Alzheimer’s Dis. JAD 2018, 61, 435–457. [Google Scholar] [CrossRef] [PubMed]
  193. Lovestone, S.; Boada, M.; Dubois, B.; Hull, M.; Rinne, J.O.; Huppertz, H.J.; Calero, M.; Andres, M.V.; Gomez-Carrillo, B.; Leon, T.; et al. A phase II trial of tideglusib in Alzheimer’s disease. J. Alzheimer’s Dis. JAD 2015, 45, 75–88. [Google Scholar] [CrossRef]
  194. West, T.; Hu, Y.; Verghese, P.B.; Bateman, R.J.; Braunstein, J.B.; Fogelman, I.; Budur, K.; Florian, H.; Mendonca, N.; Holtzman, D.M. Preclinical and Clinical Development of ABBV-8E12, a Humanized Anti-Tau Antibody, for Treatment of Alzheimer’s Disease and Other Tauopathies. J. Prev. Alzheimer’s Dis. 2017, 4, 236–241. [Google Scholar] [CrossRef]
  195. Lee, S.H.; Le Pichon, C.E.; Adolfsson, O.; Gafner, V.; Pihlgren, M.; Lin, H.; Solanoy, H.; Brendza, R.; Ngu, H.; Foreman, O.; et al. Antibody-Mediated Targeting of Tau In Vivo Does Not Require Effector Function and Microglial Engagement. Cell Rep. 2016, 16, 1690–1700. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Kontsekova, E.; Zilka, N.; Kovacech, B.; Novak, P.; Novak, M. First-in-man tau vaccine targeting structural determinants essential for pathological tau-tau interaction reduces tau oligomerisation and neurofibrillary degeneration in an Alzheimer’s disease model. Alzheimer’s Res. Ther. 2014, 6, 44. [Google Scholar] [CrossRef]
  197. 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. Alzheimer’s Res. Ther. 2018, 10, 108. [Google Scholar] [CrossRef]
  198. Novak, P.; Schmidt, R.; Kontsekova, E.; Zilka, N.; Kovacech, B.; Skrabana, R.; Vince-Kazmerova, Z.; Katina, S.; Fialova, L.; Prcina, M.; et al. Safety and immunogenicity of the tau vaccine AADvac1 in patients with Alzheimer’s disease: A randomised, double-blind, placebo-controlled, phase 1 trial. Lancet Neurol. 2017, 16, 123–134. [Google Scholar] [CrossRef]
  199. Fitzgerald, D.P.; Emerson, D.L.; Qian, Y.; Anwar, T.; Liewehr, D.J.; Steinberg, S.M.; Silberman, S.; Palmieri, D.; Steeg, P.S. TPI-287, a new taxane family member, reduces the brain metastatic colonization of breast cancer cells. Mol. Cancer Ther. 2012, 11, 1959–1967. [Google Scholar] [CrossRef]
  200. Etminan, M.; Gill, S.; Samii, A. Effect of non-steroidal anti-inflammatory drugs on risk of Alzheimer’s disease: Systematic review and meta-analysis of observational studies. BMJ 2003, 327, 128. [Google Scholar] [CrossRef]
  201. Breitner, J.C.; Haneuse, S.J.; Walker, R.; Dublin, S.; Crane, P.K.; Gray, S.L.; Larson, E.B. Risk of dementia and AD with prior exposure to NSAIDs in an elderly community-based cohort. Neurology 2009, 72, 1899–1905. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  202. Vlad, S.C.; Miller, D.R.; Kowall, N.W.; Felson, D.T. Protective effects of NSAIDs on the development of Alzheimer disease. Neurology 2008, 70, 1672–1677. [Google Scholar] [CrossRef] [Green Version]
  203. in t’ Veld, B.A.; Ruitenberg, A.; Hofman, A.; Launer, L.J.; van Duijn, C.M.; Stijnen, T.; Breteler, M.M.; Stricker, B.H. Nonsteroidal antiinflammatory drugs and the risk of Alzheimer’s disease. N. Engl. J. Med. 2001, 345, 1515–1521. [Google Scholar] [CrossRef] [PubMed]
  204. Daniels, M.J.; Rivers-Auty, J.; Schilling, T.; Spencer, N.G.; Watremez, W.; Fasolino, V.; Booth, S.J.; White, C.S.; Baldwin, A.G.; Freeman, S.; et al. Fenamate NSAIDs inhibit the NLRP3 inflammasome and protect against Alzheimer’s disease in rodent models. Nat. Commun. 2016, 7, 12504. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Wang, J.; Tan, L.; Wang, H.F.; Tan, C.C.; Meng, X.F.; Wang, C.; Tang, S.W.; Yu, J.T. Anti-inflammatory drugs and risk of Alzheimer’s disease: An updated systematic review and meta-analysis. J. Alzheimer’s Dis. JAD 2015, 44, 385–396. [Google Scholar] [CrossRef] [PubMed]
  206. Miguel-Alvarez, M.; Santos-Lozano, A.; Sanchis-Gomar, F.; Fiuza-Luces, C.; Pareja-Galeano, H.; Garatachea, N.; Lucia, A. Non-steroidal anti-inflammatory drugs as a treatment for Alzheimer’s disease: A systematic review and meta-analysis of treatment effect. Drugs Aging 2015, 32, 139–147. [Google Scholar] [CrossRef] [PubMed]
  207. Jaturapatporn, D.; Isaac, M.G.; McCleery, J.; Tabet, N. Aspirin, steroidal and non-steroidal anti-inflammatory drugs for the treatment of Alzheimer’s disease. Cochrane Database Syst. Rev. 2012, 2, CD006378. [Google Scholar] [CrossRef]
  208. 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]
  209. Muntimadugu, E.; Dhommati, R.; Jain, A.; Challa, V.G.; Shaheen, M.; Khan, W. Intranasal delivery of nanoparticle encapsulated tarenflurbil: A potential brain targeting strategy for Alzheimer’s disease. Eur. J. Pharm. Sci. 2016, 92, 224–234. [Google Scholar] [CrossRef]
  210. Turner, R.S.; Thomas, R.G.; Craft, S.; van Dyck, C.H.; Mintzer, J.; Reynolds, B.A.; Brewer, J.B.; Rissman, R.A.; Raman, R.; Aisen, P.S.; et al. A randomized, double-blind, placebo-controlled trial of resveratrol for Alzheimer disease. Neurology 2015, 85, 1383–1391. [Google Scholar] [CrossRef] [Green Version]
  211. Butchart, J.; Brook, L.; Hopkins, V.; Teeling, J.; Puntener, U.; Culliford, D.; Sharples, R.; Sharif, S.; McFarlane, B.; Raybould, R.; et al. Etanercept in Alzheimer disease: A randomized, placebo-controlled, double-blind, phase 2 trial. Neurology 2015, 84, 2161–2168. [Google Scholar] [CrossRef] [Green Version]
  212. Sano, M.; Bell, K.L.; Galasko, D.; Galvin, J.E.; Thomas, R.G.; van Dyck, C.H.; Aisen, P.S. A randomized, double-blind, placebo-controlled trial of simvastatin to treat Alzheimer disease. Neurology 2011, 77, 556–563. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Duffy, J.P.; Harrington, E.M.; Salituro, F.G.; Cochran, J.E.; Green, J.; Gao, H.; Bemis, G.W.; Evindar, G.; Galullo, V.P.; Ford, P.J.; et al. The Discovery of VX-745: A Novel and Selective p38alpha Kinase Inhibitor. ACS Med. Chem. Lett. 2011, 2, 758–763. [Google Scholar] [CrossRef] [PubMed]
  214. Alam, J.; Blackburn, K.; Patrick, D. Neflamapimod: Clinical Phase 2b-Ready Oral Small Molecule Inhibitor of p38alpha to Reverse Synaptic Dysfunction in Early Alzheimer’s Disease. J. Prev. Alzheimer’s Dis. 2017, 4, 273–278. [Google Scholar] [CrossRef]
  215. Burstein, A.H.; Sabbagh, M.; Andrews, R.; Valcarce, C.; Dunn, I.; Altstiel, L. Development of Azeliragon, an Oral Small Molecule Antagonist of the Receptor for Advanced Glycation Endproducts, for the Potential Slowing of Loss of Cognition in Mild Alzheimer’s Disease. J. Prev. Alzheimer’s Dis. 2018, 5, 149–154. [Google Scholar] [CrossRef]
  216. Burstein, A.H.; Grimes, I.; Galasko, D.R.; Aisen, P.S.; Sabbagh, M.; Mjalli, A.M. Effect of TTP488 in patients with mild to moderate Alzheimer’s disease. BMC Neurol. 2014, 14, 12. [Google Scholar] [CrossRef] [PubMed]
  217. Galimberti, D.; Scarpini, E. Pioglitazone for the treatment of Alzheimer’s disease. Expert Opin. Investig. Drugs 2017, 26, 97–101. [Google Scholar] [CrossRef] [PubMed]
  218. Geldmacher, D.S.; Fritsch, T.; McClendon, M.J.; Landreth, G. A randomized pilot clinical trial of the safety of pioglitazone in treatment of patients with Alzheimer disease. Arch. Neurol. 2011, 68, 45–50. [Google Scholar] [CrossRef] [PubMed]
  219. Sato, T.; Hanyu, H.; Hirao, K.; Kanetaka, H.; Sakurai, H.; Iwamoto, T. Efficacy of PPAR-gamma agonist pioglitazone in mild Alzheimer disease. Neurobiol. Aging 2011, 32, 1626–1633. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Microglia induced neuroinflammation in Alzheimer’s disease. Under the pathology of Alzheimer’s disease (AD), the accumulation of amyloid-β (Aβ) plagues and tau neurofibrillary tangles induce microglial M1-like activation, which produce inflammatory cytokines and cause neuronal cell death. Meanwhile, M2-like microglia is able to reduce Aβ plagues and tau neurofibrillary tangles accumulation by phagocytosis.
Figure 1. Microglia induced neuroinflammation in Alzheimer’s disease. Under the pathology of Alzheimer’s disease (AD), the accumulation of amyloid-β (Aβ) plagues and tau neurofibrillary tangles induce microglial M1-like activation, which produce inflammatory cytokines and cause neuronal cell death. Meanwhile, M2-like microglia is able to reduce Aβ plagues and tau neurofibrillary tangles accumulation by phagocytosis.
Ijms 20 00558 g001
Table 1. Drug development for Alzheimer’s disease (AD).
Table 1. Drug development for Alzheimer’s disease (AD).
DrugDescriptionPhaseCT IdentifierStatus
Drugs Target Amyloid β Production
Acitretinα-secretase enhancerIINCT01078168Completed
Epigallocatechin-Gallate (EGCG)α-secretase enhancer, prevent amyloid-β (Aβ) aggregationII/IIINCT00951834Completed
Etazolate (EHT-0202)γ-aminobutyric acid GABA)A receptor modulator, α-secretase enhancerIINCT00880412Completed
Lanabecestat (AZD3293, LY3314814)β-secretase inhibitorII/IIINCT02245737 NCT02972658 NCT02783573Terminated
LY3202626β-secretase inhibitorIINCT02791191Terminated
LY2286721β-secretase inhibitorI/IINCT01561430Terminated
Verubecestat (MK-8931)β-secretase inhibitorII/IIINCT01739348 NCT01953601Terminated
Atabecestat (JNJ-54861911)β-secretase inhibitorII/IIINCT02569398 NCT01760005Active, not recruiting
II/IIINCT02406027Terminated
Elenbecestat (E2609)β-secretase inhibitorIIINCT03036280 NCT02956486Recruiting
IINCT02322021Active, not recruiting
CNP520β-secretase inhibitorIINCT02565511 NCT03131453Recruiting
Semagacestatγ-secretase inhibitorIIINCT01035138 NCT00762411 NCT00594568Completed
Avagacestat (BMS-708163)γ-secretase inhibitorIINCT00890890 etc.Terminated
Drugs prevent Amyloid β Aggregation
PBT2metal protein-attenuating compound (MPAC), Aβ aggregation inhibitorII/IIINCT00471211Terminated
Scyllo-inositol (ELND005, AZD-103)inositol stereoisomer, Aβ aggregation inhibitorIINCT00934050, NCT00568776, NCT01735630Completed
Tramiprosate (3APS)Prevent β-sheet formation, Aβ aggregation inhibitorIIINCT00314912, NCT00088673, NCT00217763Unknown
GV-971Aβ aggregation inhibitorIIINCT02293915Completed
Immunotherapy
AN-1792 (AIP-001)Anti-Aβ vaccineIINCT00021723Terminated
CAD106Anti-Aβ vaccine, induce Anti-Aβ antibodyIINCT02565511Recruiting
Vanutide cridificar (ACC-001)Anti-Aβ vaccineIINCT00960531, etc.Terminated
Bapineuzumab (AAB-001)Anti-Aβ monoclonal antibodyIIINCT00676143, etc.Terminated
Solanezumab (LY2062430)Anti-Aβ IgG1 monoclonal antibodyIIINCT01127633 NCT01900665 NCT02760602Terminated
II/IIINCT02008357 NCT01760005Active, not recruiting
Ponezumab (PF-04360365)Anti-Aβ IgG2 antibodyIINCT00722046, NCT00945672Completed
GSK933776Anti-Aβ antibodyINCT00459550, NCT01424436Completed,
LY2599666Aβ antibodyINCT02614131Terminated
Octagam® 10%Immune globulin intravenous, 10% solutionIIINCT01736579 NCT01524887Terminated
II/IIINCT01561053 NCT01300728Active, not recruiting
IINCT03319810Enrolling by invitation
Aducanumab (BIIB037)Anti-Aβ IgG1 monoclonal antibodyIIINCT02484547 NCT02477800Active, not recruiting
IINCT03639987Recruiting
INCT01677572Active, not recruiting
Crenezumab (MABT5102A, RG7412)Anti-Aβ IgG4 antibodyI/II/IIINCT02670083 NCT01998841 NCT02353598Active, not recruiting
IIINCT03491150 NCT03114657Recruiting
Gantenerumab (R1450)Anti-Aβ IgG1 antibodyIIINCT01224106 NCT02051608Active, not recruiting
IIINCT03444870 NCT03443973Recruiting
II/IIINCT01760005Active, not recruiting
SAR228810Anti-Aβ monoclonal antibodyINCT01485302Completed
Drugs Target Tau Production
TRx0014Methylene blue, tau aggregation inhibitorIINCT00684944 NCT00515333Completed
LMTM (TRx0237)Methylene blue, tau aggregation inhibitorII/IIINCT03446001Recruiting
NCT03539380Available
IIINCT01689246 NCT01689233Completed
Tideglusib (NP031112)GSK3-β inhibitor, prevent tau hyperphosphorylationI/IINCT00948259 NCT01350362Completed
ABBV-8E12Anti-tau antibodyIINCT02880956Recruiting
IINCT03712787Not yet recruiting
RO 7105705Anti-tau antibodyINCT02820896Completed
IINCT03289143Recruiting
AADvac1Tau vaccineINCT02031198 NCT01850238Completed
IINCT02579252Active, not recruiting
TPI 287abeo-taxane, bind on tubulin, and stabilize microtubuleINCT01966666Active, not recruiting
Drugs Target Inflammation
Ibuprofennon-steroidal anti-inflammatory drugs (NSAIDs)IIINCT02547818Recruiting
TarenflurbilNSAIDsIIINCT00380276 NCT00322036Terminated
SalsalateNSAIDsINCT03277573Recruiting
CelecoxibNSAIDsIIINCT00007189Completed
ResveratrolPhenol, antioxidantIINCT01504854 NCT01716637 NCT00678431Completed
EtanerceptTumor necrosis factor-alpha (TNF-α) inhibitorI/IINCT01068353 NCT01716637Completed
Simvastatin3-hydroxy-3-methyl-glutaryl-coenzyme A (HMG-CoA) reductase inhibitor, cholesterol targetingIINCT00939822Active, not recruiting
Neflamapimod (VX-745)p38 mitogen-activated serine/threonine protein kinase p38 MAPK (p38 MAPKα) selective inhibitorIINCT03435861 NCT03402659Recruiting
Azeliragon (TTP488)Receptor for advanced glycation endproducts (RAGE) inhibitorIIINCT02080364 NCT02916056Terminated
PioglitazonePeroxisome-proliferator activated receptor γ (PPARγ) agonistsIIINCT01931566 NCT02284906Terminated

Share and Cite

MDPI and ACS Style

Dong, Y.; Li, X.; Cheng, J.; Hou, L. Drug Development for Alzheimer’s Disease: Microglia Induced Neuroinflammation as a Target? Int. J. Mol. Sci. 2019, 20, 558. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms20030558

AMA Style

Dong Y, Li X, Cheng J, Hou L. Drug Development for Alzheimer’s Disease: Microglia Induced Neuroinflammation as a Target? International Journal of Molecular Sciences. 2019; 20(3):558. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms20030558

Chicago/Turabian Style

Dong, Yuan, Xiaoheng Li, Jinbo Cheng, and Lin Hou. 2019. "Drug Development for Alzheimer’s Disease: Microglia Induced Neuroinflammation as a Target?" International Journal of Molecular Sciences 20, no. 3: 558. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms20030558

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