Review Open Access
Copyright ©The Author(s) 2015. Published by Baishideng Publishing Group Inc. All rights reserved.
World J Pharmacol. Sep 9, 2015; 4(3): 236-264
Published online Sep 9, 2015. doi: 10.5497/wjp.v4.i3.236
Therapeutic targets and delivery challenges for Alzheimer’s disease
Preshita Desai, Harshad Shete, Vandana Patravale, Department of Pharmaceutical Sciences and Technology, Institute of Chemical Technology, Mumbai 400019, Maharashtra, India
Rahul Adnaik, Department of Pharmacology, Rajarambapu College of Pharmacy, Sangli 415404, Maharashtra, India
John Disouza, Tatyasaheb Kore College of Pharmacy, Kolhapur 416113, Maharashtra, India
Author contributions: Desai P and Shete H contributed equally to the paper; all the authors contributed to this paper.
Conflict-of-interest statement: The authors declare that there is no conflict of interest.
Open-Access: This article is an open-access article which was selected by an in-house editor and fully peer-reviewed by external reviewers. It is distributed in accordance with the Creative Commons Attribution Non Commercial (CC BY-NC 4.0) license, which permits others to distribute, remix, adapt, build upon this work non-commercially, and license their derivative works on different terms, provided the original work is properly cited and the use is non-commercial. See: http://creativecommons.org/licenses/by-nc/4.0/
Correspondence to: Vandana Patravale, PhD, Professor of Pharmaceutics, Department of Pharmaceutical Sciences and Technology, Institute of Chemical Technology, Nathalal Parekh Marg, Matunga (East), Mumbai 400019, Maharashtra, India. vbp_muict@yahoo.co.in
Telephone: +91-22-33612217 Fax: +91-22-33611020
Received: November 23, 2014
Peer-review started: November 23, 2014
First decision: December 12, 2014
Revised: May 29, 2015
Accepted: August 10, 2015
Article in press: August 11, 2015
Published online: September 9, 2015

Abstract

Dementia, including Alzheimer’s disease, the 21st Century epidemic, is one of the most significant social and health crises which has currently afflicted nearly 44 million patients worldwide and about new 7.7 million cases are reported every year. This portrays the unmet need towards better understanding of Alzheimer’s disease pathomechanisms and related research towards more effective treatment strategies. The review thus comprehensively addresses Alzheimer’s disease pathophysiology with an insight of underlying multicascade pathway and elaborates possible therapeutic targets- particularly anti-amyloid approaches, anti-tau approaches, acetylcholinesterase inhibitors, glutamatergic system modifiers, immunotherapy, anti-inflammatory targets, antioxidants, 3-hydroxy-3-methyl-glutaryl-coenzyme A reductase inhibitors and insulin. In spite of extensive research leading to identification of newer targets and potent drugs, complete cure of Alzheimer’s disease appears to be an unreached holy grail. This can be attributed to their ineffective delivery across blood brain barrier and ultimately to the brain. With this understanding, researchers are now focusing on development of drug delivery systems to be delivered via suitable route that can circumvent blood brain barrier effectively with enhanced patient compliance. In this context, we have summarized current drug delivery strategies by oral, transdermal, intravenous, intranasal and other miscellaneous routes and have accentuated the future standpoint towards promising therapy ultimately leading to Alzheimer’s disease cure.

Key Words: Neurofibrillary tangles, Alzheimer’s disease, Dementia, Amyloid β, Tau, Neurodegeneration, Blood brain barrier, Transdermal, Nasal

Core tip: Dementia, including Alzheimer’s disease, the 21st Century epidemic, is one of the most significant social and health crises which has currently afflicted nearly 44 million patients worldwide and is on rampant rise. This portrays the unmet need towards better understanding of Alzheimer’s disease pathomechanisms and related research towards more effective treatment strategies. The review thus focuses on thorough understanding of Alzheimer’s disease pathophysiology, pharmacotherapy in terms of explored therapeutic targets and drug delivery systems towards better delivery of anti-Alzheirmer actives and a possible way ahead.



INTRODUCTION

Dementia, including Alzheimer’s disease (AD), the 21st Century epidemic, is one of the most significant social and health crises impacting families, social service and healthcare delivery systems.

The incidence of dementia and AD escalates almost exponentially with age[1]. The prevalence of dementia nearly doubles every five years after the age of 60 in which AD accounts for between 50%-70% among all dementia cases[2]. The age-standardized occurrence for those aged 60 or older is 5%-7%; among persons aged 60-64 years is 7%-18%, but among those aged over 90 years is 29%-64%[3-6]. It is generally believed that men and women are equally at risk of AD. However, there are more women patients than men possibly due to higher longevity of women as compared to men. Further, it is devastating to note that nearly one in four people with AD hide or conceal their symptoms, citing social stigma or dread of being ostracized[7] and four out of ten sufferers report being excluded from the familiar and comforting routines of everyday life[8].

Worldwide, approximately 44 million patients are reported to be afflicted with AD or other dementias and about 7.7 million new cases are reported every year[9]. The numbers are estimated to reach 76 million by 2030 and more than 135 million by 2050[4,10,11], with 90% increase in Europe, 226% in Asia, 248% in America and 345% in Africa[12]. In fact, most countries are woefully unprepared for the dementia epidemic and have not structured their health care programs to cope with the foreseen increase in numbers. Despite the urgent need for action, only 13 of the 193 World Health Organization members have instigated national dementia plans, precisely all of them in the developed world[13].

On the other hand, as per the current statistics, the number of cases of AD in Asia and Africa is lower than that reported in developed countries. There are several possible reasons like undiagnosed AD, the lack of awareness, poor access to technologically advanced health care, etc., or there may be lower incidence of risk factors[14]. Research in India and Africa proposes that the AD risk was possibly greater for urban as compared to rural areas. The reason for this difference is not clear whether it is increased life expectancy, lifestyle or diet?

AD though has a genetic predisposition in terms of mutations in specific genes (discussed in subsequent section), the expected hike in AD afflicted population can be attributed to increased exposure to AD risk factors that include ageing, oxidative stress (age and lifestyle induced), cardiovascular disorders, brain injuries, occupational hazards, etc.[12,15,16].

Further, the annual cost of AD related drug sales is reported to be increasing proportionally at growth rate of 33% from $500 million (year 1999) to approximately $6 billion (year 2008) and the estimated AD market is expected to cover a market size of $9.5 billion to $15 billion by year 2015-2017 (Figure 1)[16].

Figure 1
Figure 1 Schematic of Alzheimer’s disease afflicted patient population and associated therapy market data. AD: Alzheimer’s disease.

These huge statistical numbers clearly portray the unfulfilled need in AD therapeutic research and better management strategies. The major hurdle in this context is not only the identification of potential targets and discovery of potent therapeutic agents but also their effective delivery across brain.

With due consideration to these burning issues, the review focuses on thorough understanding of AD pathophysiology, pharmacotherapy in terms of explored therapeutic targets and current state of art in drug delivery systems towards better delivery of AD actives and a possible way ahead.

AD: PATHOLOGY AND SYMPTOMS

AD is a progressive brain disorder wherein the patients show clinical symptoms after a significant manifestation of disease which can take as long as 20 years[15,17]. The symptomatic appearance of AD results from progressive neurodegeneration resulting from alteration in normal anatomy and physiology of central nervous system (CNS). This primarily includes abnormal appearance of extracellular senile plaques and intracellular neurofibrillary tangles (NFTs) in CNS that interfere with classical neuronal activity triggering the neuronal death.

The senile plaques comprise toxic Amyloid β [Aβ(1-42)] protein fragments resulting from atypical amyloidogenic cleavage of amyloid precursor protein (APP). These Aβ fragments undergo sequential aggregation process to form insoluble senile plaques that get deposited in extracellular neuronal matrix. These plaques then interfere with synaptic signal transfer and induce stress signals that activate microglia, lysosomes and synaptic mitochondria ultimately causing neuronal death[15,18-21].

The intracellular NFTs are predominantly made up of hyperphosphorylated tau protein inter-tangles that impede neuronal nutrient supply leading to neuronal death. Additionally, other pathological variations like inflammation, activated microglias, elevated levels of proinflammatory cytokines, etc., accelerate the neuronal death.

From the site specific AD manifestation per se, the early neurodegeneration is observed in the cholinergic region of basal forebrain that results in cholinergic neuronal death. This results in acetylcholine (ACh) imbalance leading to early symptoms and memory loss via interference in both nicotinic and muscarinic receptor activities[18,19]. This early clinical stage of AD is commonly identified with mild to moderate forgetfulness in routine activities, apathy, depression, etc. These symptoms are broadly classified under a general class of dementia. An important point to note here is that forebrain region is associated with memory formation and thus early manifestation of AD leads to loss of recent memory followed by the old memory as the disease advances[15,19,20].

As the disease progresses senile plaques and NFTs deposition gets extrapolated to other regions of brain that predominantly include parietal and temporal lobes, hippocampus and entorhinal cortex[19,22-24]. This worsens the neuropsychiatric symptoms resulting in delirium, disorientation, lack of judgment, withdrawal from social appearance, difficulty in performing routine activities like eating, talking, walking, writing, etc.[15,19].

As the disease progresses, the brain shows high degree of shrinkage and debris deposition due to excessive neuronal death in all regions of brain. This impairment makes the patients dependent on help even for performing routine daily activities and this is identified as the final stage of the disease. At this stage, the excessively deprived brain function deprives the control on all the other body functions. This makes the patient highly vulnerable to secondary diseases like cardiac/pulmonary complications and out borne infections like pneumonia, etc., which forms the predominant reason for patient’s death[15].

AD: THERAPEUTIC TARGETS

From the ongoing multidirectional research on AD etiology, it is well evident that there is no unanimous opinion suggesting a single mechanistic pathway. Hence the pathophysiological and symptomatic advents associated with AD are believed to be resulting from a multicascade pathway leading to neurodegeneration. To understand this gradual and irreversible cognitive decline, various hypotheses have been proposed that include, formation of Aβ and extracellular fibrillation thereof, development of intracellular hyperphosphorylated tau and associated NFTs, oxidative stress, etc., ultimately resulting in neuronal death (Figure 2).

Figure 2
Figure 2 Schematic representation indicating interlinks among important Alzheimer’s disease pathogenetic hypotheses that majorly include amyloid cascade hypothesis, taupathy, mitochondrial dysfunction hypothesis, cholesterol induced mechanism, insulin signal downregulation, calpain-cathepsin hypothesis. APP: Amyloid precursor protein; PS: Presinilin; APoE: Apolipoprotein E; PI3K: Phosphoinositide-3-kinase; AKT: Protein kinase B; ROS: Reactive oxygen species.

An extensive research on these variable pathways has resulted in identification of multiple therapeutic targets which are summarized below.

Amyloid cascade and therapeutic targets

This hypothesis was proposed by Hardy and Higgins in early 1990’s and till date it is the most-researched and conceptual framework for AD which has markedly influenced drug development over a period of last 25 years[21]. The hypothesis is based on formation and accumulation of toxic Aβ(1-42) fragments resulting from abnormal amyloidogenic cleavage of trans membrane APP resulting from mutation in APP and presenilin gene (PS-1, PS-2) that regulate the entire pathway (familial origin)[25,26]. The so formed insoluble Aβ fragments further associate to form senile plaques, diffuse plaques, and cerebrovascular deposits which are the hallmarks of AD and being toxic they result in synaptic loss, neuronal death (predominantly cholinergic neurons) leading to progressive cognitive impairment[18,22-24].

Conventionally, 3 enzymes that play a crucial role in natural proteolytic cleavage of APP are α, β, and γ secretase. The first step herein comprises cleavage of extracellular fragment by α-secretase (non-amyloidogenic and predominant pathway under normal condition) or β-secretase (amyloidogenic pathway predominant under AD) leading to 83 or 99 amino acid peptide residues respectively that remain attached as a trans membrane fragment. Further, these fragments are invariably cleaved by γ-secretase which leads to formation of toxic Aβ(1-42) fragments in case of amyloidogenic pathway and initiates the extracellular plaque formation[22,24] (Figure 3).

Figure 3
Figure 3 Schematic representation of amyloid cascade hypothesis. Three enzymes α, β, and γ secretase play a crucial role in the proteolytic cleavage of APP. The first step, extracellular fragment of APP is cleaved by α-secretase (non-amyloidogenic and predominant pathway under normal condition) or β-secretase (amyloidogenic pathway predominant under AD) leading to 83 or 99 amino acid peptide residues respectively that remain attached as a trans membrane fragment. These fragments are ultimately cleaved by γ-secretase which leads to formation of toxic Aβ(1-42) fragments in case of amyloidogenic pathway and initiates the extracellular Aβ plaque formation. APP: Amyloid precursor protein; AD: Alzheimer’s disease.

Thus, targeting Aβ cascade presents the most important strategy towards management of AD. Several of such approaches include inflection of Aβ formation, augmentation of Aβ degradation, inhibition of Aβ assembly, and immunization (passive and active) to raise antibodies that target and remove Aβ and are discussed in subsequent sections.

α-secretase stimulators: This approach came in scientific limelight with an in vivo study that demonstrated the potential of enzyme ADAM 10 (a member of disintegrin and metalloproteinase family) that functions as α-secretase to prevent plaque formation and additionally it offered neuronal protection in hippocampal region[27]. This is attributed to the fact that α-secretase cleaves APP in a non-amyloidogenic pathway (Figure 3) and thus, up regulation of this enzyme is postulated to arrest Aβ formation. In this context, naturally occurring retinoids are reported to possess α-secretase stimulator activity and one such molecule acitretin is at phase 2 trial stage[28]. This finding suggests the use of natural retinoid rich food which includes spinach, carrots, soy products, etc., as a possible nutritional supplement for AD patients. Apart from natural sources, synthetic agonists of α-secretase are under thorough investigation and one such molecule, EHT-0202 has shown very promising results both in vitro and in vivo and is currently under 3-mo phase 2 clinical evaluation in 35 AD subjects[29-31].

β-secretase modulators: The β-secretase enzyme initiates the amyloidogenic pathway and thus it is a prime requisite to develop inhibitors of the same. The enzyme is very large structurally and poses difficulties in producing an inhibitor especially with an ability to cross the blood brain barrier (BBB). Thus, small molecules are being designed to inhibit the enzyme at the active site. CTS-21166, a β-secretase inhibitor is successfully reported to reduce plasma Aβ levels in phase 1 study conducted in 48 healthy volunteers at 6 different doses up to 225 mg and phase 2 study is planned[32]. In another study, central Aβ levels were lowered by the orally administrable non peptide molecule LY2811376 (molecule by Eli Lilly Inc.) in preclinical studies but further progress was halted as it affected animal retinal epithelium[33,34]. Other β-secretase inhibitor KMI-429 is being developed and human trial data is awaited[35]. Thus, this strategy is in its infancy and has to undergo a battery of safety and efficacy studies prior to becoming a market reality.

γ-secretase modulators:γ-secretase, the ultimate enzyme in amyloid cascade pathway, presents the next probable target to arrest amyloid cascade. With this in vision, MK-0752 (Merck), a γ-secretase inhibitor was developed which is in phase 2 trial as phase 1 trial was successful and indicated significant reduction in cerebrospinal fluid (CSF) Aβ levels in healthy volunteers[29,36].

Structurally, γ-secretase is a trans-membrane complex of four proteins: presenilin, presenilin enhancer 2, nicastrin, and anterior pharynx-defective 1[29,37] that play role in proteolysis of type-1 transmembrane proteins. Thus, it is worthy to note here that, apart from APP, γ-secretase has other substrates like Notch, E-cadherin, ErbB4, CD44, tyrosinase, alcadein which play a crucial role in embryogenesis and development[37]. Thus, non-selective inhibition of this protein may lead to side/adverse effects. As an instance, semagacestat (non-selective γ-secretase inhibitor) has advanced in therapeutic trials for AD but a phase 2 trial (14 wk) in 51 subjects (15, 22 and 14 subjects received placebo, 100 mg and 140 mg drug daily respectively) have shown high risk of skin rash and hair colour change which was reversed with treatment withdrawal[38]. Thus designing of an inhibitor to this enzyme desires meticulous selection. Owing to these observations, the new molecules are being developed with an aim to modulate the enzyme which will retain the therapeutic efficacy but overrule the adverse drug reactions[39].

Inhibitors of Aβ aggregation: Another encouraging approach for the development of novel therapeutics for treating AD is to prevent Aβ fibril formation especially by the small molecules. Neurochem Inc., a Canadian company, has developed a glycosaminoglycan mimetic Alzhemed™ which has an ability to bind to Aβ peptides and thereby inhibits the formation of Aβ aggregates. The molecule has successfully completed Phase 2 clinical trial and Phase 3 trial results are recently published wherein the data is very promising[40]. Metal ions like Cu2+ and Zn2+ are reported to augment Aβ aggregation and associated toxicity[41]. In consistency with this, a Cu/Zn chelator, clioquinol is reported to reduce CNS Aβ deposition after a 9 wk treatment in rodent model. The additional benefit of this molecule is its inherent tendency to cross BBB which is anticipated to ensure the therapeutic efficacy[42].

Aβ removal approaches: Aβ plaques are degraded by some proteases such as plasmin, neprilysin, insulin degrading enzyme, endothelin converting enzyme, angiotensin converting enzyme and metalloproteinase[43]. The levels of these Aβ degrading enzymes are observed to decline in AD and may contribute to Aβ accumulation[44]. In consistency to these observations, experimental evidence has suggested that inhibitors of plasminogen activator decrease the plasma and brain Aβ levels in transgenic animals[45] and increasing neprilysin levels through viral vector-delivered gene expression shows beneficial effects in animal models[46]. Additionally, the peptide hormone somatostatin is also reported to enhance Aβ clearance through activation of neprilysin[47]. Therefore targeting neprilysin with somatostatin or its analogs is an encouraging option in AD. This approach is quite in its infancy and demands thorough investigation.

Immunotherapy against Aβ: Immunotherapy was first explored by Schenk et al[48] for treatment of AD in a preclinical experiment involving Aβ(1-42) active immunization using PDAPP transgenic mice. So far numerous studies have shown encouraging results by both active (vaccination) and passive (monoclonal antibody) immunization. In active immunization, the Aβ peptide or fragment conjugated to a carrier protein and adjuvant which holds potential to stimulate cellular and humoral immune response is administered to the host which results in generation of anti-Aβ antibody. In passive immunization, the Aβ peptide specific antibody is directly injected into the host, thus evading the step of stimulating the host immune systems. The exact mechanism by which immunotherapy executes anti-AD activity is still not clear. However, studies conducted so far have given substantial proofs based on which few of the hypotheses are proposed viz. microglia-mediated phagocytosis, antibody mediated Aβ monomer sequestration, antibody mediated prevention of Aβ aggregation and neutralization of Aβ toxicity and antibody mediated peripheral clearance of Aβ[49-57]. Figure 4 illustrates the diagrammatic representation of the various mechanistic pathways of immunotherapy in AD.

Figure 4
Figure 4 Schematic diagram explaining different mechanistic pathways hypothesized for immunotherapy to exert Amyloid β clearance from brain and plasma. Immunotherapy exerts its activity by active and passive ways. In active immunization, the Aβ peptide or fragment is injected into the host which in turn stimulates cellular and humoral immune response to generate anti-Aβ antibody. In passive immunization, directly Aβ peptide specific antibody is injected into the host. The generated anti-Aβ antibodies provoke anti-AD activity by one of the or combination of following ways: microglia-mediated phagocytosis, antibody mediated Aβ monomer sequestration, antibody mediated prevention of Aβ aggregation and neutralization of Aβ toxicity and antibody mediated peripheral clearance of Aβ. Aβ: Amyloid β; BBB: Blood brain barrier. AD: Alzheimer’s disease.

Bard et al[49] and Hartman et al[50] administered Aβ monoclonal antibody to PDAPP transgenic mice and further noticed significant immunoreactivity within the microglia and macrophages. The study clearly indicated that the generated antibodies were able to cross the BBB and bind to Aβ plaques, provoking the Fc receptors (FcR)-mediated microglial phagocytosis. Numerous studies conducted by active and passive immunization have suggested that the (FcR)-mediated microglial phagocytosis might play a crucial role in clearing Aβ load from brain[49,50].

In yet another study, Yamada et al[51] found that administration of certain anti-Aβ monoclonal antibody m266 has selectively sequestered soluble Aβ monomers in the brain and terminated its progression to oligomers and plaques, thus circumventing associated neurotoxicity.

Another hypothesis suggests that certain anti-Aβ monoclonal antibodies have ability to by-pass the BBB and interact with Aβ oligomers and fibrils[52,53] to either disassemble or dissolve the existing plaques[54-56].

DeMattos et al[57] were the first to reveal the ability of antibodies to clear the Aβ levels from the systemic circulation. In this study, Aβ mid-region antibody (m266) which bears high affinity for soluble Aβ administered to PDAPP transgenic mice showed notable reduction in Aβ burden from the plasma. This mechanism was further confirmed by both active[58,59] and passive immunization[60-62]. These mechanisms indicated that anti-Aβ antibodies directly interact with plasma Aβ and enhance its clearance. This in turn imbalance the plasma to brain Aβ ratio and there by hasten the Aβ removal from the brain.

Despite the promising outcome in in vitro and preclinical studies, application of immunotherapy in clinical trial using synthetic Aβ peptide AN1792/QS-21 in AD patients with mild to moderate severity has turned out to be fatal and resulted in abrupt termination as 8 out of 300 patients developed meningoencephalitis during the study[63]. The study concluded that, there was significant plaque reduction in patients treated with AN1792/QS-2 but unfortunately it augmented T cell activation leading to meningoencephalitis. A different clinical trial with Aβ immunization, indicated noteworthy clearance of amyloid plaque in AD patients but lacked the ability to arrest neurodegenerative progression[64].

In an attempt to evade the potentially harmful T cell responses, emergence of advanced vaccines consisting sole antibody epitope(s) that lack T-cell reactive sites is a step further in the development of immunotherapy for AD. The existing data has shown promising effects while clinical reports are still awaited. In all, despite promising outcome of immunotherapy by curtailing the Aβ load and improvement of cognitive function, threat of adverse reactions still remains to be the unresolved issue[64].

Although several drugs have been investigated to be active at their intended targets, none have yet been proven to have significant clinical benefits. In 2011 and 2012, two negative trials of secretase inhibitors, semagacestat and avagacestat, and several negative trials of monoclonal antibodies, bapineuzumab and solanezumab were reported[64-68]. Recent studies have demonstrated that reducing Aβ in the brain is possible but that decreasing production or reducing fibrils or plaques is not clearly associated with clinical improvement and could be associated with toxicity[64-68]. These issues must be critically evaluated while development of newer therapeutic molecules against the specific targets in amyloid cascade hypothesis.

Taupathy and therapeutic targets

The second major hallmark of AD is formation of NFTs which comprise hyperphosphorylated form of tau protein. Tau is a protein that under normal physiological conditions stabilizes microtubules, allowing transport of vesicles and other products of neuronal cell bodies down the axon to the synapse. By regulating microtubule assembly, tau controls the morphology and growth of axons[69]. The protein has several phosphorylation sites, and the microtubule binding property of tau is dependent on the phosphorylation state. The phosphorylated tau binds microtubules with a lesser affinity leading to microtubule instability[70,71].

The trigger for this abnormal tau fibrillation initiates with stress signals (Aβ, mechanical damage, ROS, etc.) that activate primary microglial cells in CNS. This activation results in release of proinflammatory cytokines (TNFα and interleukins) that leads to neuronal alteration initiating tau hyperphosphorylation. This hyperphosphorylated form presents an aggregation tendency forming initial paired helical filaments that further precipitate to form NFTs. These intracellular NFTs elicit toxic effect on neurons leading to neuronal death. Upon neurodegeneration, these NFTs are released in extracellular matrix that in turn augments the microglial activity via positive feedback mechanism[72,73].

As per this hypothesis, AD pathology starts with formation of pretangles in proximal axons of the noradrenergic locus ceruleus which spreads via trans-synaptic transport to entorhinal cortex, hippocampus, and neocortex. This cascade of reactions is known as taupathy and is reported to worsen the AD in conjunction with senile plaques[72] (Figure 2).

Thus, it is well conceived that tau hyperphosphorylation, microtubule disruption and formation of NFTs play a crucial role in AD pathomechanisms. Therapeutically, several of these mechanisms can be targeted to arrest AD progression viz. inhibition of tau kinases to lower tau hperphosphorylation and associated aggregation, enhancing clearance of tau aggregates with drugs or antibodies and microtubule stabilization by enhancing phosphatase activity. Few of such approaches have progressed from preclinical to advanced clinical trials[74,75]. In this context, the main focus is now shifting on glycogen synthase kinase 3, a prime enzyme involved in tau phosphorylation. Lithium and valproate are reported to have inhibitory action on this enzyme and have shown promising results in terms of reducing taupathy and NFT formation in transgenic mice[76,77]. Further, it should be noted that tau pathology is not specific to AD, and occurs in several other disorders, including frontotemporal dementia, corticobasal degeneration, progressive supranuclear palsy, etc., and these approaches can be extrapolated in treatment thereof[72].

Recently, scientists have reported close relation between taupathy and synaptic mitochondrial dysfunction that leads to ROS augmentation. With this support, mitochondrial dysfunction cascade hypothesis is gaining wide attention in AD pathomechanisms[73].

Mitochondrial dysfunction and therapeutic targets

Mitochondria being the energy hose of the cells, decline in mitochondrial function allies itself with ageing and AD. The prime assumption of this hypothesis is based on genetic predisposition of AD that is presented by low genetic mitochondrial baseline function which is predominantly inherited from mother’s genome and this mitochondrial baseline function is inversely proportional with AD progression[73].

In addition to genetic predisposition, AD pathomechanisms are also reported to manifest mitochondrial dysfunction. Aβ is reported to be present in mitochondria and is observed to be interacting with complex II of respiratory chain, mitochondrial membrane and Hsp 60 (a mitochondrial chaperon matrix protein) leading to mitochondrial abnormalities. This not only alters the regular mitochondrial function but also causes abnormal increase in mitochondrial fission and reduced fusion that severely affects the mitochondrial morphology. This alteration in morphology is proposed to augment mitochondrial fragmentation and was confirmed using confocal and electron microscopic analysis in APP overexpressed neurons[78]. This can be additionally explained by the unwanted interaction of Aβ and NFT with dynamin-related protein 1, the protein that maintains the mammalian mitochondria. This interaction results in increased mitochondrial fragmentation, their restricted axonal transport and subsequent neurodegeneration[79].

Apart from genetic and AD associated factors, other environmental factors like heavy metal exposure, oxidative stress, insulin resistance, etc., are reported to down regulate the mitochondrial function via positive feedback pathway. Under normal scenario, mitochondrial biogenesis can take care of this external environmental burden but fails in case of patients with already declined neuronal activity like in case of AD. Thus, these environmental factors are postulated to cause additional affliction by hastening the mitochondrial dysfunction ultimately leading to progressive AD associated synaptic damage and symptomatic manifestation[73,80].

Oxidative stress not only causes mitochondrial dysfunction but also triggers Aβ deposition, tau hyperphosphorylation and oxidation of other neuronal components like lipids, proteins, nucleic acids, etc., causing neuronal damage[70]. The relationship between oxidative stress and AD suggests that oxidative stress is the key component of AD pathophysiology. In this context, use of antioxidants to reduce oxidative burden on cells holds a strong rationale[81].

Antioxidants for AD therapy

Antioxidant treatment is proposed to be a promising approach to slow down the disease progression by attenuating phospholipid peroxidation, protein and DNA oxidation[82]. Flavonoids and carotenoids, a group of ubiquitous antioxidants have also shown neuroprotective effect in several experiments[83,84]. Rutin, a flavonoid compound, protected rats from stress induced damage and streptozotocin induced neuronal inflammation[85]. Lutein, a natural carotenoid with cytoprotective effect[86]; when supplemented in combination with docosahexaenoic acid, memory scores and rate of learning improved in elderly women[87]. The spice curcumin has shown several beneficial roles (antioxidant, anti-inflammatory, and amyloid disaggregating properties) in experimental studies[88,89].

Melatonin is another antioxidant compound which is anticipated to be a potent anti-AD active. This is attributed to N-methyl-D-aspartate (NMDA) receptor modulation, inhibition of Aβ generation, formation of amyloid fibrils, attenuation of tau hyperphosphorylation, mitochondrial protection, and antiapoptotic effect[90]. Vitamin E has shown marked reduction in lipid peroxidation and plaque deposition when administered in transgenic AD rodent model but has failed to produce similar convincing results in humans[91]. Also, the combination of vitamin E with donepezil did not provide additional benefit in patients with AD or mild cognitive impairment[92]. Evidence for the protection offered by antioxidants including vitamins (E, C, and carotenoids), phytochemicals and synthetic compounds in AD is inconsistent[83].

Antioxidants though have shown positive results, their translation to clinic as a solitary AD therapy is not successful due to varied epidemiological data. However, it is worthy to note that they are emerging as nutritional supplements to decrease the incidence or to delay the progression of AD.

Acetylcholinesterase inhibitors for AD therapy

As discussed in earlier section (AD: Pathology and Symptoms), cholinergic neuronal death is the classical manifestation of AD leading to ACh imbalance and associated cognitive decline[18,19]. Thus restoration of CNS ACh levels is believed to offer early symptomatic relief.

In this context, cholinergic neuronal death restricts the opportunity to augment the ACh release from neurons and thus restoration of CNS ACh levels by arresting its degradation was thought to be a promising strategy. With this in view, Acetylcholinesterase (enzyme responsible for degradation of Ach) inhibitors (AChEIs) were the first amongst the pharmacological treatments sanctioned by the United States Food and Drug Administration (FDA) for AD. Currently four AChEIs are available in market: tacrine (Cognex®), donepezil (Aricept®), rivastigmine (Exelon®) and galantamine (Razadyne®, Reminyl®). Their therapeutic efficacy may be attributed to their ability to sustain cognitive function over a prolonged period of therapy[93,94]. Studies have also shown that these drugs can arrest neurodegeneration and thus can delay AD progression, if the therapy is initiated at earliest in patients with mild to moderate AD. Rivastigmine blocks butyrylcholinesterase, levels of which are reported to be augmented in the brain of patients with AD[95] and this may have an advantageous effect on prolonged cholinesterase inhibition ensuing disease stabilization. On the other hand, galantamine binds to the nicotinic ACh receptor sites which opens the ionic channels and improves the receptor responsiveness to Ach[96]. Tacrine is hardly prescribed nowadays due to its high dosing frequency and associated hepatotoxicity. Other adverse effects associated with AChEIs are nausea, vomiting, diarrhoea, anorexia, etc., and are observed to worsen during dose escalation[97]. Patients with bradycardia are at higher risk and should be given additional attention[98]. Gastrointestinal (GI) effects can be minimized by simultaneous administration of food and an anti-emetic. As the cholinergic neurons decline with disease progression (severe forms of AD), the AChEI treatments becomes inefficient.

The mechanism of action by which these drugs act is by arresting the breakdown of the neurotransmitter Ach via inhibition of the acetylcholinesterase enzyme. Acetylcholinesterase is also found to enhance Aβ plaque formation; therefore inhibition of this enzyme will not only provide symptomatic relief but also arrest AD progression[97].

Glutamatergic system modifiers for AD therapy

Glutamate is the major excitatory neurotransmitter in the CNS which is involved in a variety of functions, including synaptic neurotransmission, neuronal growth and development, synaptic plasticity, etc[98,99]. Glutamatergic neurotransmission is also observed to be very crucial in learning and memory[64,65] which is greatly hampered in case of AD patients. Moreover, the glutamatergic neurons are observed in the brain regions affected by AD, particularly in the neocortex, cortex and hippocampus[99-101].

Glutamate induced up regulation of the NMDA receptor augments the intracellular calcium level leading to neuronal death which is the hallmark of AD[102]. Memantine (Namenda®, Axura®, Ebixa®), a non-competitive NMDA receptor antagonist was approved by FDA in October 2003 for treatment of moderate to severe AD. It inhibits the neurodegeneration resulting from protracted glutamate release[103] but does not interfere with cognition at therapeutic doses[104]. At higher doses, clinical studies have shown the potential of drug to cause functional decline with delay in cognition but have overruled the possibility of any severe adverse effects[105,106]. Some studies also suggest that memantine may synergies the AChEIs therapeutic efficacy if given in combination. In one such study, it is proven that the combination was effective and well tolerated by majority of AD subjects without any severe side effects[107]. With such a promising data combination therapy can be envisioned as a better AD treatment regime.

Anti-inflammatory agents for AD therapy

The perception of AD being an inflammatory disease has appeared with two unique pathophysiological findings in AD patients. Firstly, the increased size and number of microglial cells in the brain and second the overexpression of complement system via amyloid plaques. Results from several studies, based on follow-up design and prescription based data have shown substantial reduction in the incidence of AD with prolonged use of non-steroidal anti-inflammatory drugs (NSAIDs)[108].

Investigation of AD brain has confirmed the incidence of inflammation at and around the sites of Aβ plaques deposits, NFTs and degenerating neurons. The aggregation and deposition of Aβ plaques tend to provoke the activation of microglia and astrocytes. This event is further accompanied by overexpression of complement system, particularly C1q, C3b, C3a, membrane attack complex (MAC), cytokines and chemokines. During the Aβ aggregation process, binding of C1q and C3b to amyloid plaques activates the complement system causing dystrophic neuritis and up regulating C3a and MAC (Figure 5). C3b further boosts the complement activation. C3a stimulates the activation of microglial cells which clears the Aβ through phagocytic mechanism. MAC causes cell lysis and results in toxic effects. The uncleared Aβ escalates the deposition and provokes the microglial cell activation and release of cytokines, chemokines and neurotoxins which cause neuronal loss and synaptic dysfunction[108].

Figure 5
Figure 5 Diagram representation elaborating various mechanistic pathways by which anti-Alzheimer’s disease activity is executed by non-steroidal anti-inflammatory drugs. The aggregation and deposition of Aβ plaques provoke the activation of microglia, astrocytes and complement system (C1q, C3b, C3a, MAC, cytokines and chemokines). The Aβ plaques which evades the clearance process forms deposit and provokes microglial cell activation and release of cytokines, chemokines and neurotoxins which consequently results in neuronal loss and synaptic dysfunction. NSAIDs are reported to act by multiple ways to elicit anti-AD activity which includes suppresion of oligomer formation, PPRAγ, COX, NFKβ, α1-ACT and non-amyloid α secretase pathway. Aβ: Amyloid β; APP: Amyloid precursor protein; MAC: Membrane attack complex; PPARγ: Peroxisome proliferator-activated receptor-γ; COX: Cycloxygenase; α1-ACT: α1-antichymotrypsin; BACE1: Beta-secretase 1.

Multiple mechanistic pathways have been cited for NSAIDs to elicit anti-AD activity. First of which is native anti-inflammatory action (Figure 5). NSAIDs mainly exert their anti-AD activity by suppressing the synthesis of inflammatory prostaglandin executed via Cyclooxygenase inhibition[109]. In support of this, Kotilinek et al[110] also recorded that the major improvement in the memory aspect of transgenic mice is associated with decreased prostaglandin E2 and is attributed to COX-2 inhibition. In another mechanism the NSAIDs are known to inhibit the nuclear transgenic factor Kβ which is up regulated in AD patients and is involved in regulation of many cellular target genes[111].

Yet in another investigation, the role of nuclear located peroxisome proliferator-activated receptor-γ (PPARγ) was unveiled where it is involved in regulation of pro-inflammatory genes associated with the pathogenesis of AD. The NSAIDs like ibuprofen and naproxen are known to stimulate PPARγ receptor and subsequently cause anti-inflammatory effect. Also, PPRAγ receptor mediated release of pro-inflammatory cytokines is associated with reduction of β-secretase expression and Aβ secretion[112].

The second proposed mechanism of anti-AD activity of NSAIDs is attributed to inhibition of amyloidogenic APP processing, Aβ formation and its aggregation. Studies conducted by Avramovich et al[113] have shown that NSAIDs in particular like indomethacin and ibuprofen stimulate non-amyloidogenic α-secretase pathway and cause marked release of neurotrophic and neuroprotective APP ectodomain in neuronal cells.

Further, effect of ibuprofen causing down regulation of α1-antichymotrypsin, a protein responsible for triggering Aβ pathogenesis proves another mechanistic pathway. They are known to exhibit a direct effect by inhibiting Aβ oligomer formation and subsequent deposition by interacting with Aβ peptide[114].

Despite generation of enormous data at cellular and preclinical level, replication of similar effects at clinical level has been still a matter of debate. Ten years of comprehensive study conducted in Canada to assess the incidence of AD has shown that use of NSAIDs is associated with reduced incidence of AD[115]. Contradictory to this, several studies conducted with other anti-inflammatory drugs like prednisone, hydroxychloroquine, COX-2 selective inhibitors (celecoxib, rofecoxib) and non-selective COX inhibitors (naproxen) have failed to show any advantageous effect[108,109].

3-hydroxy-3-methyl-glutaryl-CoA reductase inhibitors (statins) for AD therapy

The relation between cholesterol and AD was brought to notice for the first time by Sparks et al[116] and statistical studies have suggested that individuals on statins therapy (drugs used in cholesterol management) have very low incidence of AD.

It had been observed that elevated intracellular cholesterol induces the β-secretase activity leading to enhanced Aβ production. Inhibition of 3-hydroxy-3-methyl-glutaryl-CoA (HMG-CoA) reductase by statins, reverses this elevated intracellular cholesterol level (cholesterol dependent pathway) (Figure 6) and thereby blocks the Aβ formation. Another proposed mechanism by which statins act against Aβ progression is through cholesterol independent effect or pleiotropic effect that grants neuroprotection (Figure 6)[117,118]. Herein, statins enhance nitric oxide (NO) mediated anti-inflammatory activity and facilitate the systemic Aβ clearance by stimulating endothelial NO synthase[117,118]. Owing to these multi-target activities, statins are now emerging as promising pharmacological agents for AD treatment. The clinical trials conducted with statins like atorvastatin, lovastatin (for longer duration approximately 1 year) have shown beneficial effect by reducing the plasma Aβ levels but were not observed to be reproducible[117,118]. This could be attributed to study variations that include differences in cognitive test employed, experimental protocol, study duration and the stage of AD manifestation and dose. Thus, there is need for thorough investigation of this strategy towards its clinical approval.

Figure 6
Figure 6 Diagrammatic illustration of anti-amyloid activity of 3-hydroxy-3-methyl-glutaryl-CoA reductase inhibitors via cholesterol independent and cholesterol dependent mechanistic pathways. Cholesterol independent pathway: Statins cause NO mediated anti-inflammatory activity and facilitate the clearance of systemic Aβ by stimulating endothelial nitric oxide synthase. The statins with the virtue of its antioxidant effect cause reduction in lipid peroxidation which is escalated due to elevated levels of NO in AD brain. Cholesterol dependent pathway: The systemic cholesterol enters the brain in form of HDL. The astrocytes originated ApoE facilitates the uptake of extracellular free cholesterol and release into the neuronal cells via LRP. This act reduces free cholesterol and curtails the Aβ genesis whereas un-sequestered free ApoE aggravates the Aβ formation. Statins inhibits the HMG-CoA reductase and subsequently reverse the elevated intracellular cholesterol levels and blocks the Aβ formation. Also, ACAT inhibition leads to reduction in Aβ levels by an unknown mechanism. HDL: High density lipoprotein; BBB: Blood brain barrier; Aβ: Amyloid β; ApoE: Apolipoprotein E; LRP: LDL receptor-related protein; HMG-CoA reductase: 3-hydroxy-3-methyl-glutaryl-CoA; ER: Endoplasmic reticulum; NO: Nitric oxide; eNOS: Endothelial nitric oxide synthase; nNOS: Neuronal nitric oxide synthase; iNOS: Inducible nitric oxide synthase; ACAT: Acetyl-coenzyme a acetyltransferase. AD: Alzheimer’s disease.
Insulin for AD therapy

Apart from diabetes, insulin is reported to play a critical role in glucose uptake and neurotransmission across the brain[119-122]. Further, the proteins important for transmission of insulin signal were identified to be up regulated in AD sensitive brain regions viz. hippocampus and temporal lobe. This instigated an interest of researchers to investigate possible correlation between insulin and AD[123,124]. Remarkably, research groups both at cell-culture and clinical level have shown that normal insulin signaling under an AD condition confers beneficial effects like protection against Aβ oligomer-mediated insulin receptor loss and synaptic deterioration[125], boosting Aβ trafficking at cell membrane and clearance thereof[126].

Briefly, insulin signaling pathway is initiated when tyrosine phosphorylated insulin receptor substrate interacts with insulin receptor. Upon interaction, phosphatidylinositol 4,5-bisphosphate (PIP2) is transformed to phosphatidylinositol 3,4,5-triphosphate (PIP3) at the plasma membrane via cascade of mechanisms that finally activate Phosphoinositide-dependent kinase-1 which triggers the insulin signal. Though the exact mechanism by which insulin enhances cognitive function is unknown, the possible mechanisms can be down regulation of insulin receptors and other signaling intermediates discussed above[124-127].

Moreover, evidences also show that Aβ acts as a competitive inhibitor of insulin at the insulin receptor. This inhibition results GSK-3 stimulation by negative feedback mechanism which subsequently increases tau phosphorylation (Figure 7). Thus, insulin therapy is expected to arrest both Aβ generation and tau hyperphosphorylation[124-127].

Figure 7
Figure 7 Schematic representation explaining the role of insulin in Alzheimer’s disease treatment. Interaction of IRS with insulin receptor initiates the event of insulin signaling. This act triggers a cascade of activities which subsequently results in expression of PIP3 on the plasma membrane. This further leads to the activation of PDK-1 which interacts with AKT to generate the transmission of the insulin signal. In AD, the crucial components for insulin signaling like insulin receptors, IRS, PI3K and AKT; are down regulated. Also, Aβ competes with insulin to bind to insulin receptors, inhibits the PI3K-AKT interaction and enhances the GSK-3 activity which subsequently increases tau phosphorylation leading to impaired signaling in AD. IRS: Insulin receptor substrate; PI3K: Phosphoinositide-3-kinase; PIP: Phosphatidylinositol phosphate; PDK1: Phosphoinositide-dependent kinase-1; AKT: Protein kinase B; GSK-3: Glycogen synthase kinase-3; AD: Alzheimer’s disease.

Two dose clinical study performed in young, cognitively normal subjects (1.5 mU/kg per minute and 15 mU/kg per minute) showed improved memory performance and attention at high serum levels of insulin[127]. Other additional studies conducted in elderly impaired individuals by Craft et al[128,129] demonstrated improvement in declarative memory at dose level of 1.0 mU/kg per minute infusion. Since treatment with insulin infusion is associated with hypoglycemia, direct delivery of insulin to the brain proves to be a viable approach. In this context, intranasal administration of insulin has gained wide attention wherein a 8 wk therapy (40 IU/dose, 4 × per day) showed good performance in recalling the words in young, cognitively normal subjects[130] whereas 21 d (20 IU, 2 × per day) treatment promoted story recall and attention in cognitively impaired subjects and individuals with AD[130-132]. These effects have shown insulin to be a promising anti-AD agent.

In addition to aforementioned targets some new hypotheses are emerging with advances in understanding of AD pathmechanisms. One such recent observation is presence of autophagic vacuoles in degenerating neurons. It is proposed that these vacuoles formation results from lysosomal autophagy induced by oxidative stress, Aβ, calcium ion imbalance and cleavage of heat shock protein (Hsp 70.1) that plays a crucial role in lysosomal integrity. Accumulation of these vacuoles further results in autophagy induced neurodegeneration[133]. Thus lysosomal stabilization can be seen as a near future approach to treat AD. Thus to summarize, multivariate targets have been identified as treatment avenues for AD and are currently in various stages of clinical development. However, it must be noted that clinicians have now started believing in multi-target approach as a better therapy module to address both disease progression and symptomatic relief.

AD: DRUG DELIVERY SYSTEMS

With emergence of newer therapeutic targets as discussed in earlier section, there comes a great hope for successful AD treatment modalities in near future. Though very lucrative, the prime challenge in getting these potent drugs from bench to bedside lies in their effective delivery across BBB and ultimately to the brain. BBB is identified as an obstructive interface between blood and CNS that restricts the entry of variety of molecules to the brain via tight junctions and also serves as CNS microvasculature. This overprotective phenomenon of BBB turns out to be the rate limiting step in effective transport of drug to the brain. Further, during AD progression, BBB undergoes certain pathophysiological changes in terms of altered expression of certain transporter receptors, altered glucose transport, impaired P-gp efflux system, leaky vasculature, release of neurotoxins and oxidative stress induced changes in BBB permeability. These are being studied extensively in recent years for better understanding of disease pathophysiology. In coming years, understanding of these will play a crucial role in designing smart delivery systems to surpass BBB[134,135].

Besides drug delivery system, the route of administration also plays a significant role in drug absorption, distribution, and passage across BBB, etc. With this understanding, researchers are now focusing on development of drug delivery systems and suitable route of delivery that can synergistically circumvent BBB and enhance patient compliance[134-136]. This section describes design and development of plethora of drug delivery systems for AD therapeutics with a special attention on route of delivery. Table 1 abridges the novel strategies under investigation and Table 2 summarizes currently marketed products and ongoing clinical studies.

Table 1 Novel therapeutic strategies for Alzheimer's disease management.
ActiveMechanism of actionDrug delivery systemEfficacy study modelRef.
Oral
RutinAntioxidantOnly drugPreclinical rodent streptozotocin induced AD model[85]
(-)-Epigallocatechin-3-gallateAntioxidant, α-secretase activatorNanolipid carriersIn vitro, preclinical rodent model[137]
ChitosanPreclinical rodent model[138]
MifepristoneAntiprogesterone activity, AD symptomatic reliefPLGA nanoparticlesPreclinical rodent model[139-141]
EstradiolEstrogenic activity, AD symptomatic reliefTween 80 coated PLGA nanoparticlesPreclinical rodent model[142,143]
CNTF P021 (Ac-DGGLAG-NH2)Neurotrophic factorOnly drugPreclinical rodent AD transgenic model[144]
ClioquinolMetal ion Cu/Zn chelatorOnly drugPreclinical rodent AD transgenic model[145,146]
GalactoseGlucose restorationOnly drugPreclinical streptozotocin induced rodent model[147]
Transdermal
GalantamineAChEIDrug in adhesive type patchPreclinical rodent model[151]
Donepezil (base and salt form)AChEIFatty acid based topical formulationIn vitro skin model, preclinical rodent model[152]
Huperzine AAChEIMicroemulsion, solid lipid nanoparticles, nanostructured lipid carriersIn vitro skin model, preclinical rodent model[153]
DonepezilAChEIIontophoresis, Wearable Electronic Drug Delivery System patchesPreclinical rodent model[154]
Rasageline and selegilineMAO-B inhibitorsSolution and carbopol based gel - iontophoresisIn vitro skin model[155]
MemantineNMDA receptor modulatorIontophoresis with penetration enhancersIn vitro skin model[156]
(1-42) antigenAD immunovaccineMicroneedlesPreclinical rodent model[161]
Intravenous
ClioquinolCu/Zn chelatorOnly drugPreclinical rodent AD transgenic model[42]
LithiumGSK-3 inhibitor of tau phosphorylationOnly drugPreclinical rodent AD transgenic model[76,77]
Valproate sodiumGSK-3 inhibitor of tau phosphorylationOnly drugPreclinical rodent AD transgenic model[76,77]
Nerve growth factorCholinergic neuron protectionPolysorbate 80 coated PBCA nanoparticlesPreclinical scopolamine induced rodent model[162]
PEG chemical conjugatePreclinical rodent model[163]
Antitransferrin antibody chemical conjugatePreclinical rodent model[163,164]
GalantamineAChEIPeptide targeting ligand functionalized liposomesIn vitro cell line[165]
Aβ binding peptide QSHAβ reductionTargeted PEGylated polylactic acid nanocarriersPreclinical AD induced rodent model[166]
Thiflavin TSpecific Aβ plaque bindingPBCA NanoparticlesPreclinical rodent model[168,169]
Anti- Aβ antibodyAD immunotherapyOnly drugPreclinical rodent AD transgenic model[49,50]
BAM -10 antibodyAD immunotherapyExternal targeting with trans-cranial application of magnetic and ultrasound energyPreclinical rodent AD transgenic model[186,187]
Cholic acidCholinergic managementFluorescent labeled PEG-PLGA nanoparticles, external ultrasoundPreclinical rodent model[188]
Octapeptide NAP derived from the neurotrophic factorNeuroprotective activityOnly drugPreclinical rodent model[196,197]
Mesenchymal stem CellsNeuronal growthOnly drugIn vitro, Preclinical transgenic AD rodent model[203,204]
Intranasal
InsulinMulticascade anti AD activityOnly drugPreclinical rodent model[130-132]
TacrineAChEICyclodextrin coated bovine serum albumin nanospeheresEx vivo permeation model[175]
CurcuminAntioxidant, anti- Aβ activityNanoemulsionEx vivo permeation model[177]
siRNAAD associated gene silencingCell penetrating peptide TAT conjugated polycaprolactone- PEG MicellesEarly development[184]
Gene deliveryAD associated gene silencingExosomesEarly development[185]
Pituitary adenylate cyclase-activating polypeptideNeuroprotective activityOnly drugPreclinical rodent model[198]
Implants
RivastigmineAChEIBTM and SAM based organogel implants, subcutaneous administrationPreclinical rodent model[195]
Table 2 Current market status and ongoing clinical investigation for Alzheimer's disease therapeutics.
ActiveMechanism of actionDrug delivery routeClinical statusRef.
Approved drugs
Tacrine (Cognex®)AChEIOralUSFDA approved[205]
Donepezil (Aricept®)AChEIOralUSFDA approved
Galantamine (Razadyne, Reminyl®)AChEIOralUSFDA approved
Rivastigmine (Exelon®)AChEITransdermal patchUSFDA approved
Memantine (Namenda, Axura®, Ebixa®)NMDA receptor inhibitor, glutaminergic system modifierOralUSFDA approved
Drugs under clinical investigation
EHT-0202α-secretase activatorOralPhase 2[29]
CTS-21166β-secretase inhibitorOralPhase 2[35]
MK-0752γ-secretase inhibitorOralPhase 2[29,36]
ImmunoglobulinAD immune activityIntravenousPhase 2[170]
Omega-3-fatty acid treatment, nutritional supplement, physical exercise and cognitive stimulationAD symptomatic managementOralPhase 3[199]
Cerebrolysin and AChEINeirotropic, AChEIOralPhase 4[200,201]
Alzhemed™Activity Aβ aggregation inhibitorOral (dietary supplement)Phase 3[206]
Oral drug delivery systems

Peroral route is most preferred route of drug delivery owing to the associated patient compliance and convenience of administration. This is marked from the fact that the first approved dosage form for treatment of AD was an oral capsule of reversible AChEI drug tacrine (Cognex®).

However for AD treatment, in addition to BBB, GI stability and permeation of drug into systemic circulation become superfluous rate limiting step upon oral administration. This is well evident from the reports that Cognex® is prescribed 4 times daily which resulted in poor therapeutic compliance due to meager oral bioavailability, severe first-pass metabolism and peripheral side effects. Thus to leverage the benefits of oral delivery towards effective treatment strategy in case of AD, researchers are working towards the modified oral dosage forms that ensure better stability and permeation across GI tract.

Various drugs investigated in literature towards modified oral dosage form include AChEIs, polyphenols, metal chelators, peptides, etc. Polyphenols are a class of molecules which are extensively explored for AD treatment. One such polyphenol is (-)-epigallocatechin-3-gallate (EGCG) which is reported to be a potent activator of α-secretase that activates nonamyloidogenic processing APP but is a poor candidate for oral delivery. In this context, Smith et al[137] developed the nanolipid particles of EGCG using a co-solubilization method at drug to excipient ratio ranging from 1:1 to 1:32. The in vitro studies in murine neuroblastoma cells indicated significant enhancement in α-secretase activity above the ratio of 1:8 and activity was attributed to the better encapsulation and stability of drug in nanolipid matrix. Further, the oral pharmacokinetic studies performed in male Sprague Dawley rats exhibited over two fold enhancement in oral bioavailability with nanolipid EGCG formulation as compared to plain drug when administered at dose of 100 mg/kg of EGCG. These results signify importance of modified dosage form towards better oral bioavailability and can be anticipated to also enhance the brain uptake owing to the nanolipid matrix[137]. On similar lines, Dube et al[138] reported chitosan-tripolyphosphate nanoparticles of EGCG and performed pharmacokinetic studies in Swiss Outbred mice. The studies indicated almost 1.5 fold improvement in oral bioavailability as compared to EGCG suspension and interestingly indicated higher permeation from the jejunum region of GI tract indicating better potential of these NPs to enter systemic circulation.

Among the various AD treatment targets antiprogesterone drugs like mifepristone are also reported to arrest the cognitive impairment and thus offer symptomatic relief[139-141]. To enhance oral bioavailability, He et al[139] have reported polylactide-co-glycolide (PLGA) nanoparticles of this drug and have shown significantly high oral bioavailability of mifepristone as compared to plain drug. Thus lipid as well as polymeric encapsulation of actives presents a promising strategy towards enhanced stability and permeation of drug across GI tract. A step ahead in this field is to explore dual modalities towards enhanced GI as well as BBB permeability. For this, Mittal et al[142,143] developed estradiol entrapped PLGA nanoparticles coated with tween 80 using single emulsion technique (particle size approximately 170 nm). The rationale here was to enhance GI uptake of estradiol by nanopolymeric encapsulation and to enhance the brain uptake by tween 80 coat which is reported to be recognized and actively taken up across BBB by apolipoprotein receptors present on the BBB. Upon oral administration in Sprague Dawley rats, the estradiol level was found to be 1.969 ± 0.197 ng/g brain tissue in case of tween 80 coated nanoparticles which was almost 2 fold as compared to 1.105 ± 0.136 ng/g tissue as in case of uncoated nanoparticles. Interestingly, the drug levels obtained in brain with tween 80 coated nanoparticles given orally were almost similar to that obtained after intramuscular injection (2.123 ± 0.370 ng/g tissue) of estradiol indicating the superiority of targeted nanoparticles. Also, the pharmacodynamic studies performed using elevated plus maze test indicated marked reduction in anxiety behaviour with nanoparticulate formulation as compared to plain drug[142,143].

Oral therapy though poses issues w.r.t. low oral bioavailability, chronic oral treatment is reported to be effective probably for drugs that offer minimal or no peripheral side effects. This can be attributed to the time dependent accumulation and slow clearance from brain. In support of this, very recently Kazim et al[144] reported that the chronic oral treatment of neurotrophic factors results in reduction of neural plasticity and associated cognitive impairment. In this study, a ciliary neurotrophic factor P021 (Ac-DGGLAG-NH2) was given orally over a period of 12 mo to both moderate and severe stages of AD in transgenic mice. It was also observed that there was a significant down regulation of hyperphosphorylated tau and Aβ and thus chronic therapy can be visualized to be an effective strategy for such drugs that do not offer peripheral side effects but pose problem with oral bioavailability alone. In one such study, clioquinol (metal ion Cu/Zn chelator) was given orally over a period of 9 wk to aged APP2576 transgenic mice with advanced AD. As hypothesized, the results indicated significant reduction in both cerebral and serum Aβ levels[145,146].

Nutrient imbalance is generally observed with AD and among all, glucose hypometabolism is one such hallmark condition. With an aim to restore the brain glucose levels, Salkovic-Petrisic et al[147] supplemented streptozotocin induced AD rats with oral galactose (200 mg/kg per day) over a period of 1 mo. The pharmacodynamic evaluation with Morris Water Maze and Passive Avoidance test indicated restoration of cognitive function in galactose treated group as compared to control. Thus, it can be well noted that chronic treatment with nutrient sugars that can be converted to glucose via alternative pathways can be a supportive therapy ensuring symptomatic relief.

Transdermal drug delivery systems

Considering life-long therapy, transdermal delivery of AD actives is considered to be an ideal route as it offers sustained drug delivery over prolonged period of time with reduced dosing frequency. This certainly ensures patient compliant therapy module and positively reduces patient dependence on caretaker. Additionally, it also overrules the adversities associated with oral route viz. peripheral side effects, first pass metabolism, fluctuations in plasma drug concentration, etc.[148,149]. Owing to these lucrative advantages and feasibility of dosage form development, first AChEI drug rivastigmine transdermal patch (Exelon®) was introduced in market in year 2007 for treatment to mild to moderate AD. In a multicentric study with approximately 2000 AD patient caregivers, it was observed that 94.3% of caregivers preferred transdermal form of rivastigmine over oral therapy and corroborated better efficacy and symptomatic relief in patients[150]. With this great success of Exelon®, various conventional patches have been investigated for transdermal delivery of AD therapeutics. In one such study, Park et al[151] formulated drug in adhesive type patch of galantamine with a series of pressure sensitive adhesives. Among all, DT-2510 was found to be the most suitable pressure sensitive adhesive and this optimised patch demonstrated sustained drug plasma levels over a period of 24 h with 80% bioavailability. Thus, the successful delivery of galantamine via transdermal route proves to be a promising alternative to oral therapy ensuring relief from peripheral side effects like severe vomiting, nausea, etc.

Drug as well drug carrier both play a crucial role in delivery of drugs across the skin. To understand this phenomenon, Choi et al[152] investigated the permeation of AChEI drug donepezil in its base as well as salt form and further studied the influence of fatty acids as penetration enhancers on permeation behavior. The in vitro permeation studies performed using mouse and human cadaver skin indicated a good degree of correlation and confirmed a parabolic relationship between drug permeation and fatty acid chain length wherein the oleic acid and palmitoleic acid were observed to be optimum for the base and the salt form respectively. In vivo pharmacokinetic studies performed in rat model indicated that the base form of the drug exhibited 6 fold higher bioavailability as compared to the salt form, when formulated and applied topically with the respective fatty acid. This confirmed that the base form is more permeable and can be correlated to the lipophilic nature of drug along with the fatty acid matrix that ensured better permeation.

Considering the need of AD therapeutics to cross BBB, one must ensure higher plasma drug levels so as to felicitate BBB transport. Use of transdermal delivery presents a hurdle here, as stratum corneum, the uppermost layer of skin epidermis acts as a strong barrier towards the permeation of drug from skin into the systemic circulation. Thus higher degree and extent of drug permeation across skin demands modifications in transdermal delivery systems.

Lipid matrix based formulations are anticipated to enhance transdermal permeation and if formulated in nanocarriers are expected to cross BBB via passive transport. In context of this, Patel et al[153] developed microemulsion (ME), solid lipid nanoparticles (SLNs), and nanostructured lipid carriers (NLCs) based gels of an AChEI drug Huperzine A (particle size less than 150 nm) and compared their efficacy both in vitro and in vivo. In vitro skin permeation studies performed using rat skin resulted in highest flux observed with ME based formulation followed by NLCs and SLNs respectively. This enhanced permeation with ME based formulation can be ascribed to unique properties of ME to cause structural alteration in stratum corneum by virtue of the ME excipient matrix. Further, in vivo pharmacodynamic studies performed using elevated plus maze test in scopolamine induced mice amnesia model displayed significant reduction in transfer latency period indicating better cognition with nanoformulations as compared to orally administered drug suspension.

In order to further enhance permeation across the skin, various techniques viz. iontophoresis, sonophoresis, microneedles are being investigated. Among these techniques iontophoresis has gained the widest attention and is being explored for varied arena of AD therapeutics viz. AChEI, monoamine oxidase B (MAO-B) inhibitors, metal chelators, NMDA receptor antagonists, etc.

one such study Saluja et al[154] studied the iontophoretic delivery and effect of electric current on drug permeation via Wearable Electronic Drug Delivery patches using donepezil as a model drug. The donepezil gel loaded electronic patches were applied on hairless rats and in vivo pharmacokinetic studies were performed. The studies revealed that at current intensity of 0.13 mA, 0.26 mA and 0.39 mA, the Cmax level of drug in plasma was observed to be 0.094 μg/mL, 0.237 μg/mL and 0.336 μg/mL respectively indicating that current density has a proportional effect on drug permeation. The imperative role of electric current was also confirmed by the fact that during iontophoresis linear pharmacokinetics were observed and it altered to flip flop kinetics after iontophoretic intervention.

Kalaria et al[155] investigated anodal iontophoresis as a technique to check permeation of two MAO-B inhibitor drugs rasagiline and selegiline, both from solution and carbopol gel form. In vitro studies performed using porcine and human skin revealed that rise in electric current intensity increases the permeation flux proportionally and the major mechanism of permeation was electromigration in presence of counter ions. Further, the degree of permeation from solution form was better as compared to the carbopol gel and was attributed to slower diffusion of drugs from the gel matrix. Thus, the studies suggested that simple transdermal patches can be employed for effective iontophoretic delivery wherein the patch system plays a very crucial role on rate of permeation and thus demands a meticulous selection.

With an aim to compare the potential of chemical penetration enhancers and iontophoresis, del Rio-Sancho et al[156] conducted 2 sets of in vitro skin permeation studies wherein the skin was pre-treated over a period of 12 h with various classes of chemical permeation enhancers viz. decenoic acid, R-(+)-limonene, oleic acid, cineol, laurocapram followed by in vitro permeation with memantine drug solution. In other set, iontophoresis was investigated as a drug permeation technique at the current density of 0.5 mA/cm2. Amongst the various penetration enhancers, R-(+)-limonene exhibited maximum transdermal flux of 91.9 ± 8.2 μg/cm2per hour. Iontophoresis exhibited transdermal flux of 158 ± 6 μg/cm2 per hour which was almost 1.5 fold higher than the optimized permeation enhancer. Thus, iontophoresis can be visualized as an effective technique and additionally this opens up newer doors to investigate the co-application of both the techniques towards better permeation profiles.

As discussed in earlier section, AD immunovaccines are now emerging as a different dimension to AD therapeutics. Initial studies have proven potential of Aβ antigen injections in mouse but clinical trials were withdrawn from phase 2 due to meningoencephalitis induced by TH1 cells[63,157,158]. To overcome this adverse reaction, transcutaneous immunization is now gaining attention in recent years as it involves immune response via Th2 pathway[159,160]. Matsuo et al[161] amalgamated the concept of transcutaneous immunization with novel transdermal devices and have developed Aβ(1-42) antigen incorporated microneedle array, MicroHyala that is dissolved upon incorporation into the skin releasing the antigen. The scientists have proven the induction of immune system activation and have suggested that further modification is desired in delivery systems to achieve higher immune response and better cognitive regain.

Intravenous drug delivery systems

Intravenous delivery of AD actives is being investigated extensively as it results in 100% bioavailability. This ensures higher systemic levels of drug that presents higher probability of drug permeation across the BBB. Further, this serves as an effective research tool to investigate the effect on drug carrier system on BBB permeability as upon intravenous administration the only major rate limiting step towards entry of drug in brain is BBB.

In recent times, with a rationale to augment the delivery of nanoparticulate carriers across BBB, use of targeting ligands is gaining wide attention. Having understood the potential of Nerve growth factor (NGF) in maintenance of cholinergic neurons, targeted delivery of NGF is under extensive exploration. Kurakhmaeva et al[162] developed polybutylcyanoacrylate (PBCA) nanoparticles of NGF followed by a polysorbate 80 coat that serves as a targeting ligand across BBB. Upon intravenous administration in scopolamine induced amnesic rodent model, significant improvement in cognition was observed and was corroborated with higher levels of NGF detected in murine brain as compared to plain NGF.

Several other approaches that involve drug ligand chemical conjugation are also reported in literature for delivery of NGF that include covalent ligation of NGF to polyethylene glycol (PEG) that warrants long circulation time in systemic circulation boosting the chances of BBB uptake[163], conjugation to antitransferrin antibody enabling receptor mediated active transport across BBB[164,165]. Mufamadi et al[165] recently reported peptide targeting ligand functionalized liposomes (size approximately 150 nm) incorporating galantamine and they have shown selective uptake of these targeted nanoliposomes across PC12 cells in contrast to the non-targeted liposomes.

Use of multiple targeting ligands is also being investigated. Zhang et al[166] employed two targeting ligands on PEGylated polylactic acid (PLA) nanoparticles. The targeting ligands were TGN, a 12 amino acid ligand specific for BBB transport and an Aβ binding peptide QSH. The hypothesis here was to achieve a better permeation across BBB via active transport of target specific nanocarriers followed by selective binding to Aβ plaques. The in vivo biodistribution studies followed by intravenous administration of these nanoparticles to AD induced mice indicated almost 1.5 fold higher uptake in cerebellum and hippocampus as compared to blood indicating the preferential uptake in brain.

In another approach, Bana et al[167] have reported development of liposomes with phosphatidic acid and a derivative of ApoE-peptide as a dual targeting strategy. The in vivo biodistribution studies in rodent model upon intravenous administration revealed a higher uptake as compared to monofunctionalised liposomes. This signposts that concurrent targeting strategy can be used as a synergistic method to enhance BBB permeability.

The nanoparticulate systems are also extrapolated for diagnostic purpose. Thioflavin T is a fluorescent marker and is reported to possess specific binding affinity towards Aβ plaques. Taking advantage of this, scientists have developed polymeric PBCA nanoparticles of this dye and have shown imaging potential of this dye towards AD diagnosis. These results are very encouraging as these nanoparticles are reported to be taken across BBB from the systemic circulation and thus the current techniques of direct CNS intervention via intracerebral or intracortical or intrahippocampal injections for diagnostic purpose can be circumvented[168,169].

Intravenous immunoglobulins are reported to contain anti-Aβ antibodies and are under clinical studies for their therapeutic assessment. In an open study, 8 patients with mild AD were subjected to intravenous immunoglobulins for a period of 6 mo followed by a break that was continued with further treatment for 9 mo. Studies publicized significant reduction in Aβ levels in CSF and a symptomatic progress was observed with 2.5 points increase in Mini-mental state[170]. In extension to this, phase 2 double blind studies in patients with mild to moderate AD suggested the efficacy of treatment and confirmed safety. However they have reported need for longer trials with higher number of patients for more clinically significant data generation[171].

Intranasal drug delivery systems

Olfactory pathway is being reconnoitered comprehensively towards the brain delivery of therapeutic actives as it is the most accessible route for circumventing BBB that allows entry via peripheral olfactory neurons and lamina propria in the CNS (Figure 8)[172-174]. In this context, both passive as well as active targeting approaches are well reported in literature. Considering the nasal epithelium and permeability of olfactory pathway, nanoparticle mediated intranasal drug delivery for treatment of AD is most widely investigated for both passive as well as active delivery modules.

Figure 8
Figure 8 Schematic representation of nose to brain uptake by olfactory pathway circumventing blood brain barrier that allows entry via peripheral olfactory neurons and lamina propria in the central nervous system. Modified from[172]. CSF: Cerebrospinal fluid.

Natural polymeric nanocarriers viz. albumin, chitosan, etc. are amongst the highly explored nanoparticles because of high degree of mucoadhesion, negligible nasal mucosa irritation and compatibility. Luppi et al[175] developed cyclodextrin coated bovine serum albumin nanospheres of tacrine (size approximately 300 nm) using coacervation technique. These nanospheres presented strong mucoadhesive properties and ex vivo permeation studies using sheep nasal mucosa indicated complete permeation within 100 min. This suggests a better possibility of brain delivery via nasal route as complete dose permeated well within the nasal mucosa clearance time (approximately 4 h).

Gao et al[176] developed 6-Coumarin, a fluorescent dye loaded lectinised nanoparticles as a tool for targeted uptake via olfactory epithelium. Briefly, Ulexeuropeus agglutinin I is reported to bind specifically to l-fucose, a lectin binding domain located on olfactory epithelium. Using this as a targeting ligand decorated on PEG-PLA nanoparticles, almost 1.7 fold enhancement in brain bioavailability was seen as compared to the non-targeted PEG-PLA nanoparticles. Further, the developed targeted nanocarrier exhibited higher affinity towards olfactory mucosa than respiratory mucosa and was attributed to nanoparticulate surface immobilization of carbohydrate binding pockets present in the nasal mucosa. This evidently ensured the selective passage of targeted nanoparticles via olfactory pathway upon nasal administration.

Researchers have also developed nanoemulsion based formulation of anti-AD drugs with a view of better permeation possibility achieved using selective surfactants. In this area, Sood et al[177] formulated nanoemulsion of curcumin using high hydrophilic lipophilic balance surfactant with size less than 100 nm (optimized using Box-Behnken model). Further, with a rationale to synergise permeation by better mucoadhesion, the formulation was additionally loaded with 1% chitosan solution. Interestingly, in ex vivo permeation studies across sheep nasal mucosa, the mucoadhesive nanoemulsion exhibited an increased flux (445.1 ± 37.48 μg/cm2 per hour) in contrast to plain nanoemulsion (359.9 ± 36.85 μg/cm2 per hour). This could be attributed to opening of tight nasal epithelial junctions with enhanced permeation resulting from the combination approach. Thus, multiple permeation pathways may enhance the brain drug delivery via nasal route.

As discussed in earlier sections, Aβ is the major culprit causing AD induced neuronal death. Since past two decades, serious efforts are being directed to use this as a diagnostic marker. In this context, it must be understood that detection of Aβ in blood is a difficult challenge as it gets assimilated in blood only after sufficient progression of disease and is present at very low concentration as compared to the brain. This makes AD diagnosis a difficult task and if at all diagnosed it is only after a significant progression of disease. Thus, it is of prime concern to identify other efficient diagnostic techniques. For this, Kameshima et al[178,179] did a systematic study in Tg2576 mice wherein they not only indicated significant amount of Aβ(1-42) in nasal cavity but also proposed that it reaches nasal cavity via non blood pathway. They further proved that there is positive correlation between nasal and brain Aβ(1-42) levels which was not observed in case of serum and CSF. This is worthy to note as using nasal cavity for AD diagnosis will not only avail early diagnosis of AD but will also enable to monitor disease progression.

The efficacy of nasal route to deliver therapeutic actives is well perceived with pilot clinical studies performed for insulin delivery in patients with AD and mild cognitive disorders. The CSF biomarkers were positively identified ascertaining delivery of insulin via nasal route. Further, the study results were promising in terms of attention and memory improvement corroborating that insulin signaling pathway plays a crucial role in cognition[180]. This favors further exploration of this route for delivery of peptides which is otherwise difficult by other routes.

With this promising milestone achieved, the research here is paving a new path towards nasal delivery of siRNA, dsDNA, miRNA, etc. These molecules are recognized to be playing a key role in gene silencing especially for the pathways that lead to Aβ generation and thus represent a newer therapy module. From AD per se miR-107, miR-206 are recognized as potential actives inhibiting β-secretase and brain-derived neurotrophic factor (BDNF) respectively, whereas miR-34 is selective in terms of reducing stored BDNF levels[181-183]. Though it is lucrative, one must understand that the biggest challenge here is their site specific delivery owing to their extremely fragile nature and lack of permeability[183]. The smart formulation approach is anticipated to come to the rescue here wherein scientists have shown improved delivery of siRNA intranasally when given via cell penetrating peptide TAT conjugated polycaprolactone-PEG micelles as carriers[183,184]. Alternatively, exosomes, specialized vesicles resulting from plasma membrane like structures are coming in limelight as delivery vehicles via nasal route after proving their better efficacy via intravenous route. But research in this direction is in quite infancy and desires thorough investigation[183,185].

Altering BBB permeability

Thorough insights of AD pathophysiology and treatment strategies conclude in one major understanding that BBB serves as a major milestone to be overcome towards effective management of AD.

In supplement to various active and passive targeting strategies to cross BBB, transitory increase in BBB permeability using magnetic resonance or ultrasound is proposed to be the most site specific and sophisticated targeting strategies as it allows external control over BBB permeability and can further be focused to the particular site in the CNS.

Using this technology, Jordão et al[186] proposed the targeted immunotherapy approach for AD treatment. In this study, anti-Aβ antibody BAM-10 was administered intravenously with magnetic resonance imaging (MRI) and focused ultrasound contrast reagents in transgenic AD mice model. This was followed by trans-cranial application of magnetic and ultrasound energy. The results indicated immediate entry of contrast agents along with the antibody across the BBB and were confirmed with significant binding of antibody to Aβ plaques in brain cortical region. The similar group further explored the potential of only ultrasound energy to achieve selective BBB permeation. In the study, they injected a single dose anti-Aβ antibody with ultrasound contrast agent in transgenic AD mice model and demonstrated a significant reduction in Aβ plaques post 4 d of treatment. They also observed that the ultrasound application resulted in activation of glial cells and astrocytes in brain region which is assumed to further reduce the Aβ plaque load[187].

This strategy can be extended to ensure synergistic penetration of nanoparticles by virtue of their size along with increased permeability of BBB. To study this phenomenon, Nance et al[188] prepared fluorescent labeled PEG-PLGA nanoparticles (size approximately 60 nm) of cholic acid and administered them intravenously with contrast reagents. Upon application of brain focal MR-guided ultrasound, a significant increase in fluorescent intensity was observed as compared to when administered without MR-guided ultrasound. In another study, Treat et al[189] have demonstrated better penetration of liposomal formulations across BBB using ultrasound treatment.

These strategies are also extrapolated for AD diagnostics[190-192]. Under current protocol, an intra cerebroventricular injection of MRI contrast agent is given for imaging based diagnosis of microscopic Aβ plaques. To avoid the CNS intervention, Santin et al[192] have reported a novel ultra sound-Gd-staining protocol wherein they propose to administer clinically approved MRI contrast agent Dotarem® and microbubbles Sonovue® intravenously followed by external ultrasound treatment that ensures partitioning of contrast agent inside BBB. With preliminary studies in mouse model, the group has demonstrated fast imaging within 30 min with a resolution upto 29 μm which is similar to the one achieved with intra cerebroventricular injection alone.

These studies reveal the potential of transient improvisation of BBB permeability towards better diagnostics and therapeutic efficacy. Though non-invasive, one must thoroughly study the effect of prolong use of such strategies on BBB. To understand this Xie et al[193] employed the ultrasound frequency of 1 MHz over temporal bone of higher animal model, i.e., pigs for a period of 30 min and tracked the permeation of using MRI and a dye evans red. Studies revealed significant retention up to 90 min post exposure but was not observed at 120 min. This indicates that enhanced BBB permeability with ultrasound is a temporary and reversible mechanism but demands detailed investigation of chronic use of such techniques as AD treatment is a life-long therapy desiring regular therapeutic intervention.

Miscellaneous

With an aim to achieve sustained delivery of AD actives over a prolonged period of time, implants are emerging as novel delivery tools in this arena[194]. Sustained release biodegradable polymeric implants of PLGA, Polylactic acid, Lecithins, organogels of safflower oil, N-stearoyl L-alanine methylester (SAM), N-behenoyl L-tyrosine methylester (BTM), etc., and hydrogels are very well explored. In this class of biodegradable implants, organogel based implants (identified as a 3-dimensional gelator network entrapping an organic phase component and drug) are gaining wide attention in AD therapeutics due to their unique properties of insitu implant formation, better entrapment and control over the release of low molecular weight polar compounds.

In congruence with this hypothesis, Bastiat et al[195] developed BTM based organogels of rivastigmine and compared its efficacy w.r.t. SAM based organogels. These organogels were injected subcutaneously in rodent model wherein they formed an implant in situ. In vivo pharmacokinetic studies herein resulted in 2.5 fold enhancement in bioavailability with BTM based organogels as compared to SAM based organogels over a period of 35 d with minimum foreign body response thus proving the potential of these oraganogels in AD treatment modality.

Not only the drug delivery systems but AD therapeutic research is also witnessing the discovery of newer molecules originating from native brain protective factors. One such example is an octapeptide NAP derived from the neurotrophic factor (molecular weight = 825 Da). Researchers have shown the significant accumulation of this octapeptide in cortex and cerebellum of brain with reduced levels of hyperphosphorylated tau and Aβ in transgenic mouse[196,197]. In extension to this, Gozes et al[197] studied the effect of this oligopeptide on memory retention and reported improved short-term spatial memory behavior in cognition impaired rat model. This finding suggests the potential use of this peptide towards both pathophysiological and symptomatic relief in AD therapy.

One such other peptide under investigation is pituitary adenylate cyclase-activating polypeptide (PACAP) which is a strong α-secretase activator and possesses neuroprotective activity. Upon continuous intranasal application Rat et al[198] demonstrated that PACAP upregulated non amyloidogenic processing of resulting in reduction of Aβ and synergistically improved the levels of brain derived neurotrophic factor. These results are promising towards development of newer therapy modalities for AD and nasal route can be anticipated as an emerging choicest route for delivery of such molecules as it will not only ensure effective delivery to brain but will also confer stability to these fragile peptides otherwise difficult to deliver by other routes.

FUTURE PROSPECTS

With thorough insight of current state of art in AD research and due consideration to few burning facts viz. clinical failure of majority of drug candidates especially those targeted towards reduction of Aβ load, lack of any new drug approval and market entry since almost a decade necessitates identification of newer AD targets and/or modified treatment strategies.

In this setting, multi-domain treatment strategies are expected to become mainstream in coming years. For instance, a Multi-domain Alzheimer Prevention Study (MAPT) is currently ongoing in France that involves combination of omega-3-fatty acid treatment, nutritional supplement, physical exercise and cognitive stimulation in patient population over the age of 70. The results have shown very positive outcome opening a newer opportunity for AD management[199]. Also a combination of neurotrophic peptide drug, Cerebrolysin (Ever Neuro Pharma Ltd.) and AChEIs have shown very promising synergistic results and a phase 4 clinical trial is on-going[200,201].

Additionally, epigenetics is recently being discussed as a key reason of AD pathomechanisms and is correlated with alterations due to methylation of DNA and/or acylation of histones. Enzymes involved in these reactions are thus drawing wide attention of researchers and identification of potential inhibitors of these driver enzymes (histone acetyl transferase, etc.) is underway. Additionally a concept of epigenetic diet that includes vitamins (B6, B12, folate, etc.) is getting streamlined as these vitamins act as essential cofactors for the enzymes that control methylation homeostasis. Thus, a supplementary therapy to the existing treatment can be seen as a next step in AD therapeutics[200].

Neurodegeneration being the major pathophysiological cause of AD, stem cells induced neuroregeneration is becoming lucrative avenue which will not only arrest disease progression but will also offer symptomatic relief. In vitro studies have shown that mesenchymal stem cells (MSCs) augment neuronal cell differentiation, neurite growth and more importantly are resistant to taupathy[202]. Another study has recently suggested that human placental MSCs elicit significant immunomodulatory and paracrine effects leading to marked improvement in spatial and memory functions in AD transgenic mice[203]. With such promising research leads, it is envisioned that stem cells will soon serve as a survivor to AD patients.

CONCLUSION

Literature gives a wide spectrum of possibilities towards future AD treatment but in current setting complete AD cure appears to be an unreached holy grail. As a fact, 413 AD clinical trials have been conducted from year 2002 to 2012 and only 0.4% of trials have shown positive results[196]. This portrays the unmet need towards more better and deeper understanding of AD pathomechanisms and related research towards more effective treatment strategies. In this context, it is predicted that even if one succeeds in achieving two-year delay in both the onset and progression of AD, it will possibly reduce both AD prevalence and the last stage of disease by more than 20% and 30% respectively. This will in turn reduce the individual, social, and economic burden of the disease[197]. Thus the great challenge of coming decades will be to find financial and humanitarian resources towards better management of those afflicted with AD and to refine and redouble research efforts.

With more awareness and worldwide programs like AD Neuroimaging Initiative, Alzheimer’s Drug Therapy Initiative we are optimistic that the collaborative streamlined research will soon come up with a promising therapy for AD treatment and ultimately the cure.

ACKNOWLEDGMENTS

Preshita Desai is thankful to the INSPIRE Program, Department of Science and Technology, Government of India for providing senior research fellowship.

Footnotes

P- Reviewer: Knight DC, Orlacchio A S- Editor: Gong XM L- Editor: A E- Editor: Jiao XK

References
1.  Fratiglioni L, Launer LJ, Andersen K, Breteler MM, Copeland JR, Dartigues JF, Lobo A, Martinez-Lage J, Soininen H, Hofman A. Incidence of dementia and major subtypes in Europe: A collaborative study of population-based cohorts. Neurologic Diseases in the Elderly Research Group. Neurology. 2000;54:S10-S15.  [PubMed]  [DOI]  [Cited in This Article: ]
2.  Lobo A, Launer LJ, Fratiglioni L, Andersen K, Di Carlo A, Breteler MM, Copeland JR, Dartigues JF, Jagger C, Martinez-Lage J. Prevalence of dementia and major subtypes in Europe: A collaborative study of population-based cohorts. Neurologic Diseases in the Elderly Research Group. Neurology. 2000;54:S4-S9.  [PubMed]  [DOI]  [Cited in This Article: ]
3.  Prince M, Jackson J. World Alzheimer Report 2009. [Accessed 2014 Oct 20].  Available from: http://www.alz.co.uk/research/files/WorldAlzheimerReport.pdf.  [PubMed]  [DOI]  [Cited in This Article: ]
4.  Prince M, Bryce R, Albanese E, Wimo A, Ribeiro W, Ferri CP. The global prevalence of dementia: a systematic review and metaanalysis. Alzheimers Dement. 2013;9:63-75.e2.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2862]  [Cited by in F6Publishing: 2962]  [Article Influence: 269.3]  [Reference Citation Analysis (0)]
5.  Qiu C, Kivipelto M, von Strauss E. Epidemiology of Alzheimer’s disease: occurrence, determinants, and strategies toward intervention. Dialogues Clin Neurosci. 2009;11:111-128.  [PubMed]  [DOI]  [Cited in This Article: ]
6.  Prevalence of dementia in Europe. [Accessed 2014 Oct 20].  Available from: http://www.alzheimer-europe.org/Research/European-Collaboration-on-Dementia/Prevalence-of-dementia/Prevalence-of-dementia-in-Europe.  [PubMed]  [DOI]  [Cited in This Article: ]
7.  Ferri CP, Prince M, Brayne C, Brodaty H, Fratiglioni L, Ganguli M, Hall K, Hasegawa K, Hendrie H, Huang Y. Global prevalence of dementia: a Delphi consensus study. Lancet. 2005;366:2112-2117.  [PubMed]  [DOI]  [Cited in This Article: ]
8.  The global voice on Dementia - Demnetia statistics. [Accessed 2014 Oct 20].  Available from: http://www.alz.co.uk/research/statistics.  [PubMed]  [DOI]  [Cited in This Article: ]
9.  Dementia a public health Priority. [Accessed 2014 Oct 20].  Available from: http://apps.who.int/iris/bitstream/10665/75263/1/9789241564458_eng.pdf?ua=.  [PubMed]  [DOI]  [Cited in This Article: ]
10.  The Global Voice on Dementia - landmark report reveals massive global cost of alzheimer’s: 1% of global gdp and growing. [Accessed 2014 Oct 20].  Available from: http://www.alz.co.uk/media/nr100921.html.  [PubMed]  [DOI]  [Cited in This Article: ]
11.  Prince M, Guerchet M, Prina M. Policy brief for heads of government the global impact of dementia 2013-2050. [Accessed 2014 Oct 20].  Available from: http://www.alz.co.uk/research/GlobalImpactDementia2013.pdf.  [PubMed]  [DOI]  [Cited in This Article: ]
12.  Batsch NL, Mittelman MS, Blankman K, Ellenbogen M, Graham N, Hogg L, Mittler P, Piot P, Splaine M, Vernooij-Dassen M. World Alzheimer report 2012 overcoming the stigma of dementia. [Accessed 2014 Oct 20].  Available from: http://www.alz.co.uk/research/WorldAlzheimerReport2012.pdf.  [PubMed]  [DOI]  [Cited in This Article: ]
13.  The global voice on Dementia - Government Alzheimer plans. [Accessed 2014 Oct 20].  Available from: http://www.alz.co.uk/alzheimer-plans.  [PubMed]  [DOI]  [Cited in This Article: ]
14.  Prince M, Albanese E, Guerchet M, Prina M. World Alzheimer Report 2014: Dementia and Risk Reduction. [Accessed 2014 Oct 20].  Available from: http://www.alz.co.uk/research/world-report-2014.  [PubMed]  [DOI]  [Cited in This Article: ]
15.  Alzheimer’s Association. 2014 Alzheimer’s disease facts and figures. Alzheimers Dement. 2014;10:e47-e92.  [PubMed]  [DOI]  [Cited in This Article: ]
16.  Blood-brain barrier technologies and global markets. [Accessed 2014 Oct 20].  Available from: http://www.bccresearch.com/market-research/pharmaceuticals/blood-brain-technologies-phm075a.html.  [PubMed]  [DOI]  [Cited in This Article: ]
17.  Villemagne VL, Burnham S, Bourgeat P, Brown B, Ellis KA, Salvado O, Szoeke C, Macaulay SL, Martins R, Maruff P. Amyloid β deposition, neurodegeneration, and cognitive decline in sporadic Alzheimer’s disease: a prospective cohort study. Lancet Neurol. 2013;12:357-367.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1343]  [Cited by in F6Publishing: 1453]  [Article Influence: 132.1]  [Reference Citation Analysis (0)]
18.  Selkoe DJ. Alzheimer’s disease: genes, proteins, and therapy. Physiol Rev. 2001;81:741-766.  [PubMed]  [DOI]  [Cited in This Article: ]
19.  Auld DS, Kornecook TJ, Bastianetto S, Quirion R. Alzheimer’s disease and the basal forebrain cholinergic system: relations to beta-amyloid peptides, cognition, and treatment strategies. Prog Neurobiol. 2002;68:209-245.  [PubMed]  [DOI]  [Cited in This Article: ]
20.  Robbins TW, Elliott R, Sahakian BJ. Neuropsychology--dementia and affective disorders. Br Med Bull. 1996;52:627-643.  [PubMed]  [DOI]  [Cited in This Article: ]
21.  Golde TE, Schneider LS, Koo EH. Anti-aβ therapeutics in Alzheimer’s disease: the need for a paradigm shift. Neuron. 2011;69:203-213.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 308]  [Cited by in F6Publishing: 293]  [Article Influence: 22.5]  [Reference Citation Analysis (0)]
22.  Morgan C, Colombres M, Nuñez MT, Inestrosa NC. Structure and function of amyloid in Alzheimer’s disease. Prog Neurobiol. 2004;74:323-349.  [PubMed]  [DOI]  [Cited in This Article: ]
23.  Haass C, Selkoe DJ. Soluble protein oligomers in neurodegeneration: lessons from the Alzheimer’s amyloid beta-peptide. Nat Rev Mol Cell Biol. 2007;8:101-112.  [PubMed]  [DOI]  [Cited in This Article: ]
24.  Drachman DA. The amyloid hypothesis, time to move on: Amyloid is the downstream result, not cause, of Alzheimer’s disease. Alzheimers Dement. 2014;10:372-380.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 158]  [Cited by in F6Publishing: 168]  [Article Influence: 16.8]  [Reference Citation Analysis (0)]
25.  Tiiman A, Palumaa P, Tõugu V. The missing link in the amyloid cascade of Alzheimer’s disease - metal ions. Neurochem Int. 2013;62:367-378.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 58]  [Cited by in F6Publishing: 60]  [Article Influence: 5.5]  [Reference Citation Analysis (0)]
26.  Korczyn AD. The amyloid cascade hypothesis. Alzheimers Dement. 2008;4:176-178.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 74]  [Cited by in F6Publishing: 77]  [Article Influence: 4.8]  [Reference Citation Analysis (0)]
27.  Fahrenholz F, Postina R. Alpha-secretase activation--an approach to Alzheimer’s disease therapy. Neurodegener Dis. 2006;3:255-261.  [PubMed]  [DOI]  [Cited in This Article: ]
28.  Etchamendy N, Enderlin V, Marighetto A, Pallet V, Higueret P, Jaffard R. Vitamin A deficiency and relational memory deficit in adult mice: relationships with changes in brain retinoid signalling. Behav Brain Res. 2003;145:37-49.  [PubMed]  [DOI]  [Cited in This Article: ]
29.  Panza F, Solfrizzi V, Frisardi V, Capurso C, D’Introno A, Colacicco AM, Vendemiale G, Capurso A, Imbimbo BP. Disease-modifying approach to the treatment of Alzheimer’s disease: from alpha-secretase activators to gamma-secretase inhibitors and modulators. Drugs Aging. 2009;26:537-555.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 65]  [Cited by in F6Publishing: 69]  [Article Influence: 4.9]  [Reference Citation Analysis (0)]
30.  Marcade M, Bourdin J, Loiseau N, Peillon H, Rayer A, Drouin D, Schweighoffer F, Désiré L. Etazolate, a neuroprotective drug linking GABA(A) receptor pharmacology to amyloid precursor protein processing. J Neurochem. 2008;106:392-404.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 113]  [Cited by in F6Publishing: 116]  [Article Influence: 7.3]  [Reference Citation Analysis (0)]
31.  Vellas B, Sol O, Snyder PJ, Ousset PJ, Haddad R, Maurin M,  Lemarié JC, Désiré L, Pando MP. EHT0202/002 study group. EHT0202 in Alzheimer’s disease: a 3-month, randomized, placebo-controlled, double-blind study. Curr Alzheimer Res. 2011;8:203-212.  [PubMed]  [DOI]  [Cited in This Article: ]
32.  Ghosh AK, Brindisi M, Tang J. Developing β-secretase inhibitors for treatment of Alzheimer’s disease. J Neurochem. 2012;120 Suppl 1:71-83.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 191]  [Cited by in F6Publishing: 202]  [Article Influence: 15.5]  [Reference Citation Analysis (0)]
33.  May PC, Dean RA, Lowe SL, Martenyi F, Sheehan SM, Boggs LN, Monk SA, Mathes BM, Mergott DJ, Watson BM. Robust central reduction of amyloid-β in humans with an orally available, non-peptidic β-secretase inhibitor. J Neurosci. 2011;31:16507-16516.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 294]  [Cited by in F6Publishing: 288]  [Article Influence: 22.2]  [Reference Citation Analysis (0)]
34.  Cai J, Qi X, Kociok N, Skosyrski S, Emilio A, Ruan Q, Han S, Liu L, Chen Z, Bowes Rickman C. β-Secretase (BACE1) inhibition causes retinal pathology by vascular dysregulation and accumulation of age pigment. EMBO Mol Med. 2012;4:980-991.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 106]  [Cited by in F6Publishing: 115]  [Article Influence: 9.6]  [Reference Citation Analysis (0)]
35.  Vassar R, Kandalepas PC. The β-secretase enzyme BACE1 as a therapeutic target for Alzheimer’s disease. Alzheimers Res Ther. 2011;3:20.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 89]  [Cited by in F6Publishing: 97]  [Article Influence: 7.5]  [Reference Citation Analysis (0)]
36.  Rosen LB, Stone JA, Plump A, Yuan J, Harrison T, Flynn M, Dallob A, Matthews C, Stevenson D, Schmidt D. The gamma secretase inhibitor MK-0752 acutely and significantly reduces CSF Abeta40 concentrations in humans. Alzheimers Dement. 2006;2:S79.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 28]  [Cited by in F6Publishing: 29]  [Article Influence: 1.6]  [Reference Citation Analysis (0)]
37.  Zhang X, Li Y, Xu H, Zhang YW. The γ-secretase complex: from structure to function. Front Cell Neurosci. 2014;8:427.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 67]  [Cited by in F6Publishing: 91]  [Article Influence: 9.1]  [Reference Citation Analysis (0)]
38.  Fleisher AS, Raman R, Siemers ER, Becerra L, Clark CM, Dean RA, Farlow MR, Galvin JE, Peskind ER, Quinn JF. Phase 2 safety trial targeting amyloid beta production with a gamma-secretase inhibitor in Alzheimer disease. Arch Neurol. 2008;65:1031-1038.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 224]  [Cited by in F6Publishing: 245]  [Article Influence: 15.3]  [Reference Citation Analysis (0)]
39.  Kukar TL, Ladd TB, Bann MA, Fraering PC, Narlawar R, Maharvi GM, Healy B, Chapman R, Welzel AT, Price RW. Substrate-targeting gamma-secretase modulators. Nature. 2008;453:925-929.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 226]  [Cited by in F6Publishing: 245]  [Article Influence: 15.3]  [Reference Citation Analysis (0)]
40.  Kukar TL. Alzhemed™. [Accessed 2014 Oct 20].  Available from: http://www.alzforum.org/therapeutics/alzhemedtm.  [PubMed]  [DOI]  [Cited in This Article: ]
41.  Atwood CS, Moir RD, Huang X, Scarpa RC, Bacarra NM, Romano DM, Hartshorn MA, Tanzi RE, Bush AI. Dramatic aggregation of Alzheimer abeta by Cu(II) is induced by conditions representing physiological acidosis. J Biol Chem. 1998;273:12817-12826.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 796]  [Cited by in F6Publishing: 772]  [Article Influence: 29.7]  [Reference Citation Analysis (0)]
42.  Cherny RA, Atwood CS, Xilinas ME, Gray DN, Jones WD, McLean CA, Barnham KJ, Volitakis I, Fraser FW, Kim Y. Treatment with a copper-zinc chelator markedly and rapidly inhibits beta-amyloid accumulation in Alzheimer’s disease transgenic mice. Neuron. 2001;30:665-676.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1095]  [Cited by in F6Publishing: 1059]  [Article Influence: 46.0]  [Reference Citation Analysis (0)]
43.  Nalivaeva NN, Beckett C, Belyaev ND, Turner AJ. Are amyloid-degrading enzymes viable therapeutic targets in Alzheimer’s disease? J Neurochem. 2012;120 Suppl 1:167-185.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 133]  [Cited by in F6Publishing: 139]  [Article Influence: 10.7]  [Reference Citation Analysis (0)]
44.  Yasojima K, McGeer EG, McGeer PL. Relationship between beta amyloid peptide generating molecules and neprilysin in Alzheimer disease and normal brain. Brain Res. 2001;919:115-121.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 142]  [Cited by in F6Publishing: 167]  [Article Influence: 7.3]  [Reference Citation Analysis (0)]
45.  Jacobsen JS, Comery TA, Martone RL, Elokdah H, Crandall DL, Oganesian A, Aschmies S, Kirksey Y, Gonzales C, Xu J. Enhanced clearance of Abeta in brain by sustaining the plasmin proteolysis cascade. Proc Natl Acad Sci USA. 2008;105:8754-8759.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 96]  [Cited by in F6Publishing: 104]  [Article Influence: 6.5]  [Reference Citation Analysis (0)]
46.  Marr RA, Rockenstein E, Mukherjee A, Kindy MS, Hersh LB, Gage FH, Verma IM, Masliah E. Neprilysin gene transfer reduces human amyloid pathology in transgenic mice. J Neurosci. 2003;23:1992-1996.  [PubMed]  [DOI]  [Cited in This Article: ]
47.  Saito T, Iwata N, Tsubuki S, Takaki Y, Takano J, Huang SM, Suemoto T, Higuchi M, Saido TC. Somatostatin regulates brain amyloid beta peptide Abeta42 through modulation of proteolytic degradation. Nat Med. 2005;11:434-439.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 266]  [Cited by in F6Publishing: 277]  [Article Influence: 14.6]  [Reference Citation Analysis (0)]
48.  Schenk D, Barbour R, Dunn W, Gordon G, Grajeda H, Guido T, Hu K, Huang J, Johnson-Wood K, Khan K. Immunization with amyloid-beta attenuates Alzheimer-disease-like pathology in the PDAPP mouse. Nature. 1999;400:173-177.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2308]  [Cited by in F6Publishing: 2201]  [Article Influence: 88.0]  [Reference Citation Analysis (0)]
49.  Bard F, Barbour R, Cannon C, Carretto R, Fox M, Games D, Guido T, Hoenow K, Hu K, Johnson-Wood K. Epitope and isotype specificities of antibodies to beta -amyloid peptide for protection against Alzheimer’s disease-like neuropathology. Proc Natl Acad Sci USA. 2003;100:2023-2028.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 302]  [Cited by in F6Publishing: 311]  [Article Influence: 14.8]  [Reference Citation Analysis (0)]
50.  Hartman RE, Izumi Y, Bales KR, Paul SM, Wozniak DF, Holtzman DM. Treatment with an amyloid-beta antibody ameliorates plaque load, learning deficits, and hippocampal long-term potentiation in a mouse model of Alzheimer’s disease. J Neurosci. 2005;25:6213-6220.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 90]  [Cited by in F6Publishing: 110]  [Article Influence: 6.1]  [Reference Citation Analysis (0)]
51.  Yamada K, Yabuki C, Seubert P, Schenk D, Hori Y, Ohtsuki S, Terasaki T, Hashimoto T, Iwatsubo T. Abeta immunotherapy: intracerebral sequestration of Abeta by an anti-Abeta monoclonal antibody 266 with high affinity to soluble Abeta. J Neurosci. 2009;29:11393-11398.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 81]  [Cited by in F6Publishing: 92]  [Article Influence: 6.1]  [Reference Citation Analysis (0)]
52.  Solomon B, Koppel R, Hanan E, Katzav T. Monoclonal antibodies inhibit in vitro fibrillar aggregation of the Alzheimer beta-amyloid peptide. Proc Natl Acad Sci USA. 1996;93:452-455.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 218]  [Cited by in F6Publishing: 220]  [Article Influence: 7.9]  [Reference Citation Analysis (0)]
53.  Legleiter J, Czilli DL, Gitter B, DeMattos RB, Holtzman DM, Kowalewski T. Effect of different anti-Abeta antibodies on Abeta fibrillogenesis as assessed by atomic force microscopy. J Mol Biol. 2004;335:997-1006.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 79]  [Cited by in F6Publishing: 86]  [Article Influence: 4.3]  [Reference Citation Analysis (0)]
54.  Solomon B, Koppel R, Frankel D, Hanan-Aharon E. Disaggregation of Alzheimer beta-amyloid by site-directed mAb. Proc Natl Acad Sci USA. 1997;94:4109-4112.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 310]  [Cited by in F6Publishing: 315]  [Article Influence: 11.7]  [Reference Citation Analysis (0)]
55.  Frenkel D, Katz O, Solomon B. Immunization against Alzheimer’s beta -amyloid plaques via EFRH phage administration. Proc Natl Acad Sci USA. 2000;97:11455-11459.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 122]  [Cited by in F6Publishing: 134]  [Article Influence: 5.6]  [Reference Citation Analysis (0)]
56.  Bacskai BJ, Kajdasz ST, Christie RH, Carter C, Games D, Seubert P, Schenk D, Hyman BT. Imaging of amyloid-beta deposits in brains of living mice permits direct observation of clearance of plaques with immunotherapy. Nat Med. 2001;7:369-372.  [PubMed]  [DOI]  [Cited in This Article: ]
57.  DeMattos RB, Bales KR, Cummins DJ, Dodart JC, Paul SM, Holtzman DM. Peripheral anti-A beta antibody alters CNS and plasma A beta clearance and decreases brain A beta burden in a mouse model of Alzheimer’s disease. Proc Natl Acad Sci USA. 2001;98:8850-8855.  [PubMed]  [DOI]  [Cited in This Article: ]
58.  Sigurdsson EM, Wisniewski T, Frangione B. A safer vaccine for Alzheimer’s disease? Neurobiol Aging. 2002;23:1001-1008.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 37]  [Cited by in F6Publishing: 43]  [Article Influence: 2.0]  [Reference Citation Analysis (0)]
59.  Lemere CA, Spooner ET, LaFrancois J, Malester B, Mori C, Leverone JF, Matsuoka Y, Taylor JW, DeMattos RB, Holtzman DM. Evidence for peripheral clearance of cerebral Abeta protein following chronic, active Abeta immunization in PSAPP mice. Neurobiol Dis. 2003;14:10-18.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 98]  [Cited by in F6Publishing: 119]  [Article Influence: 5.7]  [Reference Citation Analysis (0)]
60.  Deane R, Du Yan S, Submamaryan RK, LaRue B, Jovanovic S, Hogg E, Welch D, Manness L, Lin C, Yu J. RAGE mediates amyloid-beta peptide transport across the blood-brain barrier and accumulation in brain. Nat Med. 2003;9:907-913.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 980]  [Cited by in F6Publishing: 1045]  [Article Influence: 49.8]  [Reference Citation Analysis (0)]
61.  Deane R, Sagare A, Hamm K, Parisi M, LaRue B, Guo H, Wu Z, Holtzman DM, Zlokovic BV. IgG-assisted age-dependent clearance of Alzheimer’s amyloid beta peptide by the blood-brain barrier neonatal Fc receptor. J Neurosci. 2005;25:11495-11503.  [PubMed]  [DOI]  [Cited in This Article: ]
62.  Dodel R, Hampel H, Depboylu C, Lin S, Gao F, Schock S, Jäckel S, Wei X, Buerger K, Höft C. Human antibodies against amyloid beta peptide: a potential treatment for Alzheimer’s disease. Ann Neurol. 2002;52:253-256.  [PubMed]  [DOI]  [Cited in This Article: ]
63.  Orgogozo JM, Gilman S, Dartigues JF, Laurent B, Puel M, Kirby LC, Jouanny P, Dubois B, Eisner L, Flitman S. Subacute meningoencephalitis in a subset of patients with AD after Abeta42 immunization. Neurology. 2003;61:46-54.  [PubMed]  [DOI]  [Cited in This Article: ]
64.  Holmes C, Boche D, Wilkinson D, Yadegarfar G, Hopkins V, Bayer A, Jones RW, Bullock R, Love S, Neal JW. 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.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1017]  [Cited by in F6Publishing: 961]  [Article Influence: 60.1]  [Reference Citation Analysis (0)]
65.  Rinne JO, Brooks DJ, Rossor MN, Fox NC, Bullock R, Klunk WE, Mathis CA, Blennow K, Barakos J, Okello AA, Rodriguez Martinez de Liano S, Liu E, Koller M, Gregg KM, Schenk D, Black R, Grundman M. 11C-PiB PET assessment of change in fibrillar amyloid-beta load in patients with Alzheimer’s disease treated with bapineuzumab: a phase 2, double-blind, placebo-controlled, ascending-dose study. Lancet Neurol. 2010;9:363-372.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 546]  [Cited by in F6Publishing: 515]  [Article Influence: 36.8]  [Reference Citation Analysis (0)]
66.  Gilman S, Koller M, Black RS, Jenkins L, Griffith SG, Fox NC, Eisner L, Kirby L, Rovira MB, Forette F, Orgogozo JM. AN1792(QS-21)-201 Study Team. Clinical effects of Abeta immunization (AN1792) in patients with AD in an interrupted trial. Neurology. 2005;64:1553-1562.  [PubMed]  [DOI]  [Cited in This Article: ]
67.  Salloway S, Sperling R, Gilman S, Fox NC, Blennow K, Raskind M, Sabbagh M, Honig LS, Doody R, van Dyck CH. A phase 2 multiple ascending dose trial of bapineuzumab in mild to moderate Alzheimer disease. Neurology. 2009;73:2061-2070.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 492]  [Cited by in F6Publishing: 517]  [Article Influence: 34.5]  [Reference Citation Analysis (0)]
68.  Ostrowitzki S, Deptula D, Thurfjell L, Barkhof F, Bohrmann B, Brooks DJ, Klunk WE, Ashford E, Yoo K, Xu ZX. Mechanism of amyloid removal in patients with Alzheimer disease treated with gantenerumab. Arch Neurol. 2012;69:198-207.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 269]  [Cited by in F6Publishing: 274]  [Article Influence: 21.1]  [Reference Citation Analysis (0)]
69.  Buée L, Bussière T, Buée-Scherrer V, Delacourte A, Hof PR. Tau protein isoforms, phosphorylation and role in neurodegenerative disorders. Brain Res Brain Res Rev. 2000;33:95-130.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1376]  [Cited by in F6Publishing: 1380]  [Article Influence: 57.5]  [Reference Citation Analysis (0)]
70.  Maccioni RB, Cambiazo V. Role of microtubule-associated proteins in the control of microtubule assembly. Physiol Rev. 1995;75:835-864.  [PubMed]  [DOI]  [Cited in This Article: ]
71.  Garcia ML, Cleveland DW. Going new places using an old MAP: tau, microtubules and human neurodegenerative disease. Curr Opin Cell Biol. 2001;13:41-48.  [PubMed]  [DOI]  [Cited in This Article: ]
72.  Maccioni RB, Farías G, Morales I, Navarrete L. The revitalized tau hypothesis on Alzheimer’s disease. Arch Med Res. 2010;41:226-231.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 166]  [Cited by in F6Publishing: 183]  [Article Influence: 13.1]  [Reference Citation Analysis (0)]
73.  Eckert A, Nisbet R, Grimm A, Götz J. March separate, strike together--role of phosphorylated TAU in mitochondrial dysfunction in Alzheimer’s disease. Biochim Biophys Acta. 2014;1842:1258-1266.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 71]  [Cited by in F6Publishing: 78]  [Article Influence: 7.1]  [Reference Citation Analysis (0)]
74.  Boutajangout A, Sigurdsson EM, Krishnamurthy PK. Tau as a therapeutic target for Alzheimer’s disease. Curr Alzheimer Res. 2011;8:666-677.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 44]  [Cited by in F6Publishing: 47]  [Article Influence: 3.9]  [Reference Citation Analysis (0)]
75.  Fuentes P, Catalan J. A clinical perspective: anti tau’s treatment in Alzheimer’s disease. Curr Alzheimer Res. 2011;8:686-688.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 5]  [Cited by in F6Publishing: 6]  [Article Influence: 0.5]  [Reference Citation Analysis (0)]
76.  Engel T, Goñi-Oliver P, Lucas JJ, Avila J, Hernández F. Chronic lithium administration to FTDP-17 tau and GSK-3beta overexpressing mice prevents tau hyperphosphorylation and neurofibrillary tangle formation, but pre-formed neurofibrillary tangles do not revert. J Neurochem. 2006;99:1445-1455.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 150]  [Cited by in F6Publishing: 161]  [Article Influence: 8.9]  [Reference Citation Analysis (0)]
77.  Noble W, Planel E, Zehr C, Olm V, Meyerson J, Suleman F, Gaynor K, Wang L, LaFrancois J, Feinstein B. Inhibition of glycogen synthase kinase-3 by lithium correlates with reduced tauopathy and degeneration in vivo. Proc Natl Acad Sci USA. 2005;102:6990-6995.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 475]  [Cited by in F6Publishing: 496]  [Article Influence: 26.1]  [Reference Citation Analysis (0)]
78.  Wang X, Su B, Siedlak SL, Moreira PI, Fujioka H, Wang Y, Casadesus G, Zhu X. Amyloid-beta overproduction causes abnormal mitochondrial dynamics via differential modulation of mitochondrial fission/fusion proteins. Proc Natl Acad Sci USA. 2008;105:19318-19323.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 565]  [Cited by in F6Publishing: 636]  [Article Influence: 39.8]  [Reference Citation Analysis (0)]
79.  Manczak M, Reddy PH. Abnormal interaction between the mitochondrial fission protein Drp1 and hyperphosphorylated tau in Alzheimer’s disease neurons: implications for mitochondrial dysfunction and neuronal damage. Hum Mol Genet. 2012;21:2538-2547.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 260]  [Cited by in F6Publishing: 309]  [Article Influence: 25.8]  [Reference Citation Analysis (0)]
80.  Picone P, Nuzzo D, Caruana L, Scafidi V, Di Carlo M. Mitochondrial dysfunction: different routes to Alzheimer’s disease therapy. Oxid Med Cell Longev. 2014;2014:780179.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 126]  [Cited by in F6Publishing: 134]  [Article Influence: 13.4]  [Reference Citation Analysis (0)]
81.  Zhu X, Lee HG, Perry G, Smith MA. Alzheimer disease, the two-hit hypothesis: an update. Biochim Biophys Acta. 2007;1772:494-502.  [PubMed]  [DOI]  [Cited in This Article: ]
82.  Markesbery WR. The role of oxidative stress in Alzheimer disease. Arch Neurol. 1999;56:1449-1452.  [PubMed]  [DOI]  [Cited in This Article: ]
83.  Kelsey NA, Wilkins HM, Linseman DA. Nutraceutical antioxidants as novel neuroprotective agents. Molecules. 2010;15:7792-7814.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 318]  [Cited by in F6Publishing: 294]  [Article Influence: 21.0]  [Reference Citation Analysis (0)]
84.  Khalili M, Hamzeh F. Effects of active constituents of Crocus sativus L., crocin on streptozocin-induced model of sporadic Alzheimer’s disease in male rats. Iran Biomed J. 2010;14:59-65.  [PubMed]  [DOI]  [Cited in This Article: ]
85.  Javed H, Khan MM, Ahmad A, Vaibhav K, Ahmad ME, Khan A, Ashafaq M, Islam F, Siddiqui MS, Safhi MM. Rutin prevents cognitive impairments by ameliorating oxidative stress and neuroinflammation in rat model of sporadic dementia of Alzheimer type. Neuroscience. 2012;210:340-352.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 206]  [Cited by in F6Publishing: 208]  [Article Influence: 17.3]  [Reference Citation Analysis (0)]
86.  Vijayapadma V, Ramyaa P, Pavithra D, Krishnasamy R. Protective effect of lutein against benzo(a)pyrene-induced oxidative stress in human erythrocytes. Toxicol Ind Health. 2014;30:284-293.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 19]  [Cited by in F6Publishing: 21]  [Article Influence: 1.8]  [Reference Citation Analysis (0)]
87.  Johnson EJ. A possible role for lutein and zeaxanthin in cognitive function in the elderly. Am J Clin Nutr. 2012;96:1161S-1165S.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 136]  [Cited by in F6Publishing: 120]  [Article Influence: 10.0]  [Reference Citation Analysis (0)]
88.  Lim GP, Chu T, Yang F, Beech W, Frautschy SA, Cole GM. The curry spice curcumin reduces oxidative damage and amyloid pathology in an Alzheimer transgenic mouse. J Neurosci. 2001;21:8370-8377.  [PubMed]  [DOI]  [Cited in This Article: ]
89.  Park SY, Kim DS. Discovery of natural products from Curcuma longa that protect cells from beta-amyloid insult: a drug discovery effort against Alzheimer’s disease. J Nat Prod. 2002;65:1227-1231.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 195]  [Cited by in F6Publishing: 172]  [Article Influence: 7.8]  [Reference Citation Analysis (0)]
90.  Wang JZ, Wang ZF. Role of melatonin in Alzheimer-like neurodegeneration. Acta Pharmacol Sin. 2006;27:41-49.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 93]  [Cited by in F6Publishing: 93]  [Article Influence: 5.2]  [Reference Citation Analysis (0)]
91.  Farina N, Isaac MG, Clark AR, Rusted J, Tabet N. Vitamin E for Alzheimer’s dementia and mild cognitive impairment. Cochrane Database Syst Rev. 2012;11:CD002854.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 44]  [Cited by in F6Publishing: 80]  [Article Influence: 6.7]  [Reference Citation Analysis (0)]
92.  Petersen RC, Thomas RG, Grundman M, Bennett D, Doody R, Ferris S, Galasko D, Jin S, Kaye J, Levey A. Vitamin E and donepezil for the treatment of mild cognitive impairment. N Engl J Med. 2005;352:2379-2388.  [PubMed]  [DOI]  [Cited in This Article: ]
93.  Mohs RC, Doody RS, Morris JC, Ieni JR, Rogers SL, Perdomo CA, Pratt RD. “312” Study Group. A 1-year, placebo-controlled preservation of function survival study of donepezil in AD patients. Neurology. 2001;57:481-488.  [PubMed]  [DOI]  [Cited in This Article: ]
94.  Raskind MA, Peskind ER, Truyen L, Kershaw P, Damaraju CV. The cognitive benefits of galantamine are sustained for at least 36 months: a long-term extension trial. Arch Neurol. 2004;61:252-256.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 106]  [Cited by in F6Publishing: 118]  [Article Influence: 5.9]  [Reference Citation Analysis (0)]
95.  Perry EK, Atack JR, Perry RH, Hardy JA, Dodd PR, Edwardson JA, Blessed G, Tomlinson BE, Fairbairn AF. Intralaminar neurochemical distributions in human midtemporal cortex: comparison between Alzheimer’s disease and the normal. J Neurochem. 1984;42:1402-1410.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 75]  [Cited by in F6Publishing: 77]  [Article Influence: 1.9]  [Reference Citation Analysis (0)]
96.  Albuquerque EX, Alkondon M, Pereira EF, Castro NG, Schrattenholz A, Barbosa CT, Bonfante-Cabarcas R, Aracava Y, Eisenberg HM, Maelicke A. Properties of neuronal nicotinic acetylcholine receptors: pharmacological characterization and modulation of synaptic function. J Pharmacol Exp Ther. 1997;280:1117-1136.  [PubMed]  [DOI]  [Cited in This Article: ]
97.  Ellis JM. Cholinesterase inhibitors in the treatment of dementia. J Am Osteopath Assoc. 2005;105:145-158.  [PubMed]  [DOI]  [Cited in This Article: ]
98.  Butterfield DA, Pocernich CB. The glutamatergic system and Alzheimer’s disease: therapeutic implications. CNS Drugs. 2003;17:641-652.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 183]  [Cited by in F6Publishing: 182]  [Article Influence: 8.7]  [Reference Citation Analysis (0)]
99.  Francis PT. Glutamatergic systems in Alzheimer’s disease. Int J Geriatr Psychiatry. 2003;18:S15-S21.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 156]  [Cited by in F6Publishing: 155]  [Article Influence: 7.4]  [Reference Citation Analysis (0)]
100.  Kowall NW, Beal MF. Glutamate-, glutaminase-, and taurine-immunoreactive neurons develop neurofibrillary tangles in Alzheimer’s disease. Ann Neurol. 1991;29:162-167.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 63]  [Cited by in F6Publishing: 66]  [Article Influence: 2.0]  [Reference Citation Analysis (0)]
101.  Bussière T, Giannakopoulos P, Bouras C, Perl DP, Morrison JH, Hof PR. Progressive degeneration of nonphosphorylated neurofilament protein-enriched pyramidal neurons predicts cognitive impairment in Alzheimer’s disease: stereologic analysis of prefrontal cortex area 9. J Comp Neurol. 2003;463:281-302.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 135]  [Cited by in F6Publishing: 131]  [Article Influence: 6.2]  [Reference Citation Analysis (0)]
102.  Lipton SA, Rosenberg PA. Excitatory amino acids as a final common pathway for neurologic disorders. N Engl J Med. 1994;330:613-622.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1974]  [Cited by in F6Publishing: 1917]  [Article Influence: 63.9]  [Reference Citation Analysis (0)]
103.  Danysz W, Parsons CG, Mobius HJ, Stoffler A, Quack G. Neuroprotective and symptomatological action of memantine relevant for Alzheimer’s disease--a unified glutamatergic hypothesis on the mechanism of action. Neurotox Res. 2000;2:85-97.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 155]  [Cited by in F6Publishing: 149]  [Article Influence: 6.2]  [Reference Citation Analysis (0)]
104.  Smith PF. Therapeutic N-methyl-D-aspartate receptor antagonists: will reality meet expectation? Curr Opin Investig Drugs. 2003;4:826-832.  [PubMed]  [DOI]  [Cited in This Article: ]
105.  Reisberg B, Doody R, Stöffler A, Schmitt F, Ferris S, Möbius HJ; Memantine Study Group. Memantine in moderate-to-severe Alzheimer’s disease. N Engl J Med. 2003;348:1333-1341.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1315]  [Cited by in F6Publishing: 1182]  [Article Influence: 56.3]  [Reference Citation Analysis (0)]
106.  Winblad B, Poritis N. Memantine in severe dementia: results of the 9M-Best Study (Benefit and efficacy in severely demented patients during treatment with memantine). Int J Geriatr Psychiatry. 1999;14:135-146.  [PubMed]  [DOI]  [Cited in This Article: ]
107.  Tariot PN, Farlow MR, Grossberg GT, Graham SM, McDonald S, Gergel I; Memantine Study Group. Memantine treatment in patients with moderate to severe Alzheimer disease already receiving donepezil: a randomized controlled trial. JAMA. 2004;291:317-324.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 911]  [Cited by in F6Publishing: 987]  [Article Influence: 49.4]  [Reference Citation Analysis (0)]
108.  Szekely CA, Thorne JE, Zandi PP, Ek M, Messias E, Breitner JC, Goodman SN. Nonsteroidal anti-inflammatory drugs for the prevention of Alzheimer’s disease: a systematic review. Neuroepidemiology. 2004;23:159-169.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 263]  [Cited by in F6Publishing: 257]  [Article Influence: 12.9]  [Reference Citation Analysis (0)]
109.  Hoozemans JJ, Rozemuller JM, van Haastert ES, Veerhuis R, Eikelenboom P. Cyclooxygenase-1 and -2 in the different stages of Alzheimer’s disease pathology. Curr Pharm Des. 2008;14:1419-1427.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 81]  [Cited by in F6Publishing: 92]  [Article Influence: 5.8]  [Reference Citation Analysis (0)]
110.  Kotilinek LA, Westerman MA, Wang Q, Panizzon K, Lim GP, Simonyi A, Lesne S, Falinska A, Younkin LH, Younkin SG. Cyclooxygenase-2 inhibition improves amyloid-beta-mediated suppression of memory and synaptic plasticity. Brain. 2008;131:651-664.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 158]  [Cited by in F6Publishing: 177]  [Article Influence: 11.1]  [Reference Citation Analysis (0)]
111.  Sung S, Yang H, Uryu K, Lee EB, Zhao L, Shineman D, Trojanowski JQ, Lee VM, Praticò D. Modulation of nuclear factor-kappa B activity by indomethacin influences A beta levels but not A beta precursor protein metabolism in a model of Alzheimer’s disease. Am J Pathol. 2004;165:2197-2206.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 114]  [Cited by in F6Publishing: 127]  [Article Influence: 6.4]  [Reference Citation Analysis (0)]
112.  Sastre M, Dewachter I, Rossner S, Bogdanovic N, Rosen E, Borghgraef P, Evert BO, Dumitrescu-Ozimek L, Thal DR, Landreth G. Nonsteroidal anti-inflammatory drugs repress beta-secretase gene promoter activity by the activation of PPARgamma. Proc Natl Acad Sci USA. 2006;103:443-448.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 264]  [Cited by in F6Publishing: 295]  [Article Influence: 16.4]  [Reference Citation Analysis (0)]
113.  Avramovich Y, Amit T, Youdim MB. Non-steroidal anti-inflammatory drugs stimulate secretion of non-amyloidogenic precursor protein. J Biol Chem. 2002;277:31466-31473.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 63]  [Cited by in F6Publishing: 68]  [Article Influence: 3.1]  [Reference Citation Analysis (0)]
114.  Morihara T, Teter B, Yang F, Lim GP, Boudinot S, Boudinot FD, Frautschy SA, Cole GM. Ibuprofen suppresses interleukin-1beta induction of pro-amyloidogenic alpha1-antichymotrypsin to ameliorate beta-amyloid (Abeta) pathology in Alzheimer’s models. Neuropsychopharmacology. 2005;30:1111-1120.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 71]  [Cited by in F6Publishing: 80]  [Article Influence: 4.2]  [Reference Citation Analysis (0)]
115.  Côté S, Carmichael PH, Verreault R, Lindsay J, Lefebvre J, Laurin D. Nonsteroidal anti-inflammatory drug use and the risk of cognitive impairment and Alzheimer’s disease. Alzheimers Dement. 2012;8:219-226.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 85]  [Cited by in F6Publishing: 88]  [Article Influence: 7.3]  [Reference Citation Analysis (0)]
116.  Sparks DL, Hunsaker JC, Scheff SW, Kryscio RJ, Henson JL, Markesbery WR. Cortical senile plaques in coronary artery disease, aging and Alzheimer’s disease. Neurobiol Aging. 1990;11:601-607.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 193]  [Cited by in F6Publishing: 203]  [Article Influence: 6.0]  [Reference Citation Analysis (0)]
117.  Cucchiara B, Kasner SE. Use of statins in CNS disorders. J Neurol Sci. 2001;187:81-89.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 90]  [Cited by in F6Publishing: 95]  [Article Influence: 4.1]  [Reference Citation Analysis (0)]
118.  Delanty N, Vaughan CJ, Sheehy N. Statins and neuroprotection. Expert Opin Investig Drugs. 2001;10:1847-1853.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 22]  [Cited by in F6Publishing: 28]  [Article Influence: 1.2]  [Reference Citation Analysis (0)]
119.  Clarke DW, Boyd FT, Kappy MS, Raizada MK. Insulin binds to specific receptors and stimulates 2-deoxy-D-glucose uptake in cultured glial cells from rat brain. J Biol Chem. 1984;259:11672-11675.  [PubMed]  [DOI]  [Cited in This Article: ]
120.  Raizada MK, Shemer J, Judkins JH, Clarke DW, Masters BA, LeRoith D. Insulin receptors in the brain: structural and physiological characterization. Neurochem Res. 1988;13:297-303.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 49]  [Cited by in F6Publishing: 53]  [Article Influence: 1.5]  [Reference Citation Analysis (0)]
121.  Smythe GA, Bradshaw JE, Nicholson MV, Grunstein HS, Storlien LH. Rapid bidirectional effects of insulin on hypothalamic noradrenergic and serotoninergic neuronal activity in the rat: role in glucose homeostasis. Endocrinology. 1985;117:1590-1597.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 40]  [Cited by in F6Publishing: 42]  [Article Influence: 1.1]  [Reference Citation Analysis (0)]
122.  Gammeltoft S, Fehlmann M, Van Obberghen E. Insulin receptors in the mammalian central nervous system: binding characteristics and subunit structure. Biochimie. 1985;67:1147-1153.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 30]  [Cited by in F6Publishing: 36]  [Article Influence: 0.9]  [Reference Citation Analysis (0)]
123.  Moloney AM, Griffin RJ, Timmons S, O’Connor R, Ravid R, O’Neill C. Defects in IGF-1 receptor, insulin receptor and IRS-1/2 in Alzheimer’s disease indicate possible resistance to IGF-1 and insulin signalling. Neurobiol Aging. 2010;31:224-243.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 515]  [Cited by in F6Publishing: 553]  [Article Influence: 39.5]  [Reference Citation Analysis (0)]
124.  Steen E, Terry BM, Rivera EJ, Cannon JL, Neely TR, Tavares R, Xu XJ, Wands JR, de la Monte SM. Impaired insulin and insulin-like growth factor expression and signaling mechanisms in Alzheimer’s disease--is this type 3 diabetes? J Alzheimers Dis. 2005;7:63-80.  [PubMed]  [DOI]  [Cited in This Article: ]
125.  De Felice FG, Vieira MN, Bomfim TR, Decker H, Velasco PT, Lambert MP, Viola KL, Zhao WQ, Ferreira ST, Klein WL. Protection of synapses against Alzheimer’s-linked toxins: insulin signaling prevents the pathogenic binding of Abeta oligomers. Proc Natl Acad Sci USA. 2009;106:1971-1976.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 465]  [Cited by in F6Publishing: 498]  [Article Influence: 33.2]  [Reference Citation Analysis (0)]
126.  Gasparini L, Gouras GK, Wang R, Gross RS, Beal MF, Greengard P, Xu H. Stimulation of beta-amyloid precursor protein trafficking by insulin reduces intraneuronal beta-amyloid and requires mitogen-activated protein kinase signaling. J Neurosci. 2001;21:2561-2570.  [PubMed]  [DOI]  [Cited in This Article: ]
127.  Kern W, Peters A, Fruehwald-Schultes B, Deininger E, Born J, Fehm HL. Improving influence of insulin on cognitive functions in humans. Neuroendocrinology. 2001;74:270-280.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 232]  [Cited by in F6Publishing: 239]  [Article Influence: 10.4]  [Reference Citation Analysis (0)]
128.  Craft S, Newcomer J, Kanne S, Dagogo-Jack S, Cryer P, Sheline Y, Luby J, Dagogo-Jack A, Alderson A. Memory improvement following induced hyperinsulinemia in Alzheimer’s disease. Neurobiol Aging. 1996;17:123-130.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 323]  [Cited by in F6Publishing: 350]  [Article Influence: 12.5]  [Reference Citation Analysis (0)]
129.  Craft S, Asthana S, Newcomer JW, Wilkinson CW, Matos IT, Baker LD, Cherrier M, Lofgreen C, Latendresse S, Petrova A. Enhancement of memory in Alzheimer disease with insulin and somatostatin, but not glucose. Arch Gen Psychiatry. 1999;56:1135-1140.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 218]  [Cited by in F6Publishing: 237]  [Article Influence: 9.5]  [Reference Citation Analysis (0)]
130.  Benedict C, Hallschmid M, Schmitz K, Schultes B, Ratter F, Fehm HL, Born J, Kern W. Intranasal insulin improves memory in humans: superiority of insulin aspart. Neuropsychopharmacology. 2007;32:239-243.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 216]  [Cited by in F6Publishing: 220]  [Article Influence: 12.9]  [Reference Citation Analysis (0)]
131.  Reger MA, Watson GS, Frey WH, Baker LD, Cholerton B, Keeling ML, Belongia DA, Fishel MA, Plymate SR, Schellenberg GD. Effects of intranasal insulin on cognition in memory-impaired older adults: modulation by APOE genotype. Neurobiol Aging. 2006;27:451-458.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 436]  [Cited by in F6Publishing: 469]  [Article Influence: 26.1]  [Reference Citation Analysis (0)]
132.  Reger MA, Watson GS, Green PS, Wilkinson CW, Baker LD, Cholerton B, Fishel MA, Plymate SR, Breitner JC, DeGroodt W. Intranasal insulin improves cognition and modulates beta-amyloid in early AD. Neurology. 2008;70:440-448.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 535]  [Cited by in F6Publishing: 543]  [Article Influence: 31.9]  [Reference Citation Analysis (0)]
133.  Yamashima T. Reconsider Alzheimer’s disease by the ‘calpain-cathepsin hypothesis’--a perspective review. Prog Neurobiol. 2013;105:1-23.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 68]  [Cited by in F6Publishing: 79]  [Article Influence: 7.2]  [Reference Citation Analysis (0)]
134.  Banks WA. Drug delivery to the brain in Alzheimer’s disease: consideration of the blood-brain barrier. Adv Drug Deliv Rev. 2012;64:629-639.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 111]  [Cited by in F6Publishing: 121]  [Article Influence: 10.1]  [Reference Citation Analysis (0)]
135.  Pahnke J, Walker LC, Scheffler K, Krohn M. Alzheimer’s disease and blood-brain barrier function-Why have anti-beta-amyloid therapies failed to prevent dementia progression? Neurosci Biobehav Rev. 2009;33:1099-1108.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 47]  [Cited by in F6Publishing: 53]  [Article Influence: 3.5]  [Reference Citation Analysis (0)]
136.  Starr JM, Farrall AJ, Armitage P, McGurn B, Wardlaw J. Blood-brain barrier permeability in Alzheimer’s disease: a case-control MRI study. Psychiatry Res. 2009;171:232-241.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 133]  [Cited by in F6Publishing: 137]  [Article Influence: 9.1]  [Reference Citation Analysis (0)]
137.  Smith A, Giunta B, Bickford PC, Fountain M, Tan J, Shytle RD. Nanolipidic particles improve the bioavailability and alpha-secretase inducing ability of epigallocatechin-3-gallate (EGCG) for the treatment of Alzheimer’s disease. Int J Pharm. 2010;389:207-212.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 182]  [Cited by in F6Publishing: 168]  [Article Influence: 12.0]  [Reference Citation Analysis (0)]
138.  Dube A, Nicolazzo JA, Larson I. Chitosan nanoparticles enhance the intestinal absorption of the green tea catechins (+)-catechin and (-)-epigallocatechin gallate. Eur J Pharm Sci. 2010;41:219-225.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 189]  [Cited by in F6Publishing: 165]  [Article Influence: 11.8]  [Reference Citation Analysis (0)]
139.  He W, Horn SW, Hussain MD. Improved bioavailability of orally administered mifepristone from PLGA nanoparticles. Int J Pharm. 2007;334:173-178.  [PubMed]  [DOI]  [Cited in This Article: ]
140.  Belanoff JK, Jurik J, Schatzberg LD, DeBattista C, Schatzberg AF. Slowing the progression of cognitive decline in Alzheimer’s disease using mifepristone. J Mol Neurosci. 2002;19:201-206.  [PubMed]  [DOI]  [Cited in This Article: ]
141.  Brambilla D, Le Droumaguet B, Nicolas J, Hashemi SH, Wu LP, Moghimi SM, Couvreur P, Andrieux K. Nanotechnologies for Alzheimer’s disease: diagnosis, therapy, and safety issues. Nanomedicine. 2011;7:521-540.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 190]  [Cited by in F6Publishing: 171]  [Article Influence: 13.2]  [Reference Citation Analysis (0)]
142.  Mittal G, Carswell H, Brett R, Currie S, Kumar MN. Development and evaluation of polymer nanoparticles for oral delivery of estradiol to rat brain in a model of Alzheimer’s pathology. J Control Release. 2011;150:220-228.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 92]  [Cited by in F6Publishing: 92]  [Article Influence: 6.6]  [Reference Citation Analysis (0)]
143.  Mittal G, Sahana DK, Bhardwaj V, Ravi Kumar MN. Estradiol loaded PLGA nanoparticles for oral administration: effect of polymer molecular weight and copolymer composition on release behavior in vitro and in vivo. J Control Release. 2007;119:77-85.  [PubMed]  [DOI]  [Cited in This Article: ]
144.  Kazim SF, Blanchard J, Dai CL, Tung YC, LaFerla FM, Iqbal IG, Iqbal K. Disease modifying effect of chronic oral treatment with a neurotrophic peptidergic compound in a triple transgenic mouse model of Alzheimer’s disease. Neurobiol Dis. 2014;71:110-130.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 57]  [Cited by in F6Publishing: 61]  [Article Influence: 6.1]  [Reference Citation Analysis (0)]
145.  Ritchie CW, Bush AI, Mackinnon A, Macfarlane S, Mastwyk M, MacGregor L, Kiers L, Cherny R, Li QX, Tammer A. Metal-protein attenuation with iodochlorhydroxyquin (clioquinol) targeting Abeta amyloid deposition and toxicity in Alzheimer disease: a pilot phase 2 clinical trial. Arch Neurol. 2003;60:1685-1691.  [PubMed]  [DOI]  [Cited in This Article: ]
146.  Roney C, Kulkarni P, Arora V, Antich P, Bonte F, Wu A, Mallikarjuana NN, Manohar S, Liang HF, Kulkarni AR. Targeted nanoparticles for drug delivery through the blood-brain barrier for Alzheimer’s disease. J Control Release. 2005;108:193-214.  [PubMed]  [DOI]  [Cited in This Article: ]
147.  Salkovic-Petrisic M, Osmanovic-Barilar J, Knezovic A, Hoyer S, Mosetter K, Reutter W. Long-term oral galactose treatment prevents cognitive deficits in male Wistar rats treated intracerebroventricularly with streptozotocin. Neuropharmacology. 2014;77:68-80.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 50]  [Cited by in F6Publishing: 55]  [Article Influence: 5.0]  [Reference Citation Analysis (0)]
148.  Oertel W, Ross JS, Eggert K, Adler G. Rationale for transdermal drug administration in Alzheimer disease. Neurology. 2007;69:S4-S9.  [PubMed]  [DOI]  [Cited in This Article: ]
149.  Utsuki T, Uchimura N, Irikura M, Moriuchi H, Holloway HW, Yu QS, Spangler EL, Mamczarz J, Ingram DK, Irie T. Preclinical investigation of the topical administration of phenserine: transdermal flux, cholinesterase inhibition, and cognitive efficacy. J Pharmacol Exp Ther. 2007;321:353-361.  [PubMed]  [DOI]  [Cited in This Article: ]
150.  Reñé R, Ricart J, Hernández B. From high doses of oral rivastigmine to transdermal rivastigmine patches: user experience and satisfaction among caregivers of patients with mild to moderate Alzheimer disease. Neurologia. 2014;29:86-93.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 11]  [Cited by in F6Publishing: 9]  [Article Influence: 0.8]  [Reference Citation Analysis (0)]
151.  Park CW, Son DD, Kim JY, Oh TO, Ha JM, Rhee YS, Park ES. Investigation of formulation factors affecting in vitro and in vivo characteristics of a galantamine transdermal system. Int J Pharm. 2012;436:32-40.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 27]  [Cited by in F6Publishing: 25]  [Article Influence: 2.1]  [Reference Citation Analysis (0)]
152.  Choi J, Choi MK, Chong S, Chung SJ, Shim CK, Kim DD. Effect of fatty acids on the transdermal delivery of donepezil: in vitro and in vivo evaluation. Int J Pharm. 2012;422:83-90.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 56]  [Cited by in F6Publishing: 60]  [Article Influence: 4.6]  [Reference Citation Analysis (0)]
153.  Patel PA, Patil SC, Kalaria DR, Kalia YN, Patravale VB. Comparative in vitro and in vivo evaluation of lipid based nanocarriers of Huperzine A. Int J Pharm. 2013;446:16-23.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 38]  [Cited by in F6Publishing: 51]  [Article Influence: 4.6]  [Reference Citation Analysis (0)]
154.  Saluja S, Kasha PC, Paturi J, Anderson C, Morris R, Banga AK. A novel electronic skin patch for delivery and pharmacokinetic evaluation of donepezil following transdermal iontophoresis. Int J Pharm. 2013;453:395-399.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 45]  [Cited by in F6Publishing: 47]  [Article Influence: 4.3]  [Reference Citation Analysis (0)]
155.  Kalaria DR, Patel P, Patravale V, Kalia YN. Comparison of the cutaneous iontophoretic delivery of rasagiline and selegiline across porcine and human skin in vitro. Int J Pharm. 2012;438:202-208.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 15]  [Cited by in F6Publishing: 16]  [Article Influence: 1.3]  [Reference Citation Analysis (0)]
156.  del Rio-Sancho S, Serna-Jiménez CE, Calatayud-Pascual MA, Balaguer-Fernández C, Femenía-Font A, Merino V, López-Castellano A. Transdermal absorption of memantin--effect of chemical enhancers, iontophoresis, and role of enhancer lipophilicity. Eur J Pharm Biopharm. 2012;82:164-170.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in F6Publishing: 1]  [Reference Citation Analysis (0)]
157.  Fox NC, Black RS, Gilman S, Rossor MN, Griffith SG, Jenkins L, Koller M. Effects of Abeta immunization (AN1792) on MRI measures of cerebral volume in Alzheimer disease. Neurology. 2005;64:1563-1572.  [PubMed]  [DOI]  [Cited in This Article: ]
158.  Bayer AJ, Bullock R, Jones RW, Wilkinson D, Paterson KR, Jenkins L, Millais SB, Donoghue S. Evaluation of the safety and immunogenicity of synthetic Abeta42 (AN1792) in patients with AD. Neurology. 2005;64:94-101.  [PubMed]  [DOI]  [Cited in This Article: ]
159.  Ishii-Katsuno R, Nakajima A, Katsuno T, Nojima J, Futai E, Sasagawa N, Yoshida T, Watanabe Y, Ishiura S. Reduction of amyloid beta-peptide accumulation in Tg2576 transgenic mice by oral vaccination. Biochem Biophys Res Commun. 2010;399:593-599.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 10]  [Cited by in F6Publishing: 13]  [Article Influence: 0.9]  [Reference Citation Analysis (0)]
160.  Nikolic WV, Bai Y, Obregon D, Hou H, Mori T, Zeng J, Ehrhart J, Shytle RD, Giunta B, Morgan D. Transcutaneous beta-amyloid immunization reduces cerebral beta-amyloid deposits without T cell infiltration and microhemorrhage. Proc Natl Acad Sci USA. 2007;104:2507-2512.  [PubMed]  [DOI]  [Cited in This Article: ]
161.  Matsuo K, Okamoto H, Kawai Y, Quan YS, Kamiyama F, Hirobe S, Okada N, Nakagawa S. Vaccine efficacy of transcutaneous immunization with amyloid β using a dissolving microneedle array in a mouse model of Alzheimer’s disease. J Neuroimmunol. 2014;266:1-11.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 38]  [Cited by in F6Publishing: 40]  [Article Influence: 3.6]  [Reference Citation Analysis (0)]
162.  Kurakhmaeva KB, Djindjikhashvili IA, Petrov VE, Balabanyan VU, Voronina TA, Trofimov SS, Kreuter J, Gelperina S, Begley D, Alyautdin RN. Brain targeting of nerve growth factor using poly(butyl cyanoacrylate) nanoparticles. J Drug Target. 2009;17:564-574.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 157]  [Cited by in F6Publishing: 140]  [Article Influence: 9.3]  [Reference Citation Analysis (0)]
163.  Brasnjevic I, Steinbusch HW, Schmitz C, Martinez-Martinez P; European NanoBioPharmaceutics Research Initiative. Delivery of peptide and protein drugs over the blood-brain barrier. Prog Neurobiol. 2009;87:212-251.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 199]  [Cited by in F6Publishing: 177]  [Article Influence: 11.8]  [Reference Citation Analysis (0)]
164.  Liao GS, Li XB, Zhang CY, Shu YY, Tang SX. Pharmacological actions of nerve growth factor-transferrin conjugate on the central nervous system. J Nat Toxins. 2001;10:291-297.  [PubMed]  [DOI]  [Cited in This Article: ]
165.  Mufamadi MS, Choonara YE, Kumar P, Modi G, Naidoo D, van Vuuren S, Ndesendo VM, Toit LC, Iyuke SE, Pillay V. Ligand-functionalized nanoliposomes for targeted delivery of galantamine. Int J Pharm. 2013;448:267-281.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 31]  [Cited by in F6Publishing: 35]  [Article Influence: 3.2]  [Reference Citation Analysis (0)]
166.  Zhang C, Wan X, Zheng X, Shao X, Liu Q, Zhang Q, Qian Y. Dual-functional nanoparticles targeting amyloid plaques in the brains of Alzheimer’s disease mice. Biomaterials. 2014;35:456-465.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 158]  [Cited by in F6Publishing: 155]  [Article Influence: 14.1]  [Reference Citation Analysis (0)]
167.  Bana L, Minniti S, Salvati E, Sesana S, Zambelli V, Cagnotto A, Orlando A, Cazzaniga E, Zwart R, Scheper W. Liposomes bi-functionalized with phosphatidic acid and an ApoE-derived peptide affect Aβ aggregation features and cross the blood-brain-barrier: implications for therapy of Alzheimer disease. Nanomedicine. 2014;10:1583-1590.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 88]  [Cited by in F6Publishing: 99]  [Article Influence: 9.0]  [Reference Citation Analysis (0)]
168.  Härtig W, Paulke BR, Varga C, Seeger J, Harkany T, Kacza J. Electron microscopic analysis of nanoparticles delivering thioflavin-T after intrahippocampal injection in mouse: implications for targeting beta-amyloid in Alzheimer’s disease. Neurosci Lett. 2003;338:174-176.  [PubMed]  [DOI]  [Cited in This Article: ]
169.  Modi G, Pillay V, Choonara YE, Ndesendo VM, du Toit LC, Naidoo D. Nanotechnological applications for the treatment of neurodegenerative disorders. Prog Neurobiol. 2009;88:272-285.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 100]  [Cited by in F6Publishing: 95]  [Article Influence: 6.3]  [Reference Citation Analysis (0)]
170.  Relkin NR, Szabo P, Adamiak B, Burgut T, Monthe C, Lent RW, Younkin S, Younkin L, Schiff R, Weksler ME. 18-Month study of intravenous immunoglobulin for treatment of mild Alzheimer disease. Neurobiol Aging. 2009;30:1728-1736.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 210]  [Cited by in F6Publishing: 246]  [Article Influence: 15.4]  [Reference Citation Analysis (0)]
171.  Dodel R, Rominger A, Bartenstein P, Barkhof F, Blennow K, Förster S, Winter Y, Bach JP, Popp J, Alferink J. Intravenous immunoglobulin for treatment of mild-to-moderate Alzheimer’s disease: a phase 2, randomised, double-blind, placebo-controlled, dose-finding trial. Lancet Neurol. 2013;12:233-243.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 134]  [Cited by in F6Publishing: 144]  [Article Influence: 13.1]  [Reference Citation Analysis (0)]
172.  Jain R, Nabar S, Dandekar P, Hassan P, Aswal V, Talmon Y, Shet T, Borde L, Ray K, Patravale V. Formulation and evaluation of novel micellar nanocarrier for nasal delivery of sumatriptan. Nanomedicine (Lond). 2010;5:575-587.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 32]  [Cited by in F6Publishing: 32]  [Article Influence: 2.3]  [Reference Citation Analysis (0)]
173.  Chiappetta DA, Hocht C, Opezzo JA, Sosnik A. Intranasal administration of antiretroviral-loaded micelles for anatomical targeting to the brain in HIV. Nanomedicine (Lond). 2013;8:223-237.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 70]  [Cited by in F6Publishing: 74]  [Article Influence: 6.2]  [Reference Citation Analysis (0)]
174.  Kumar A, Pandey AN, Jain SK. Nasal-nanotechnology: revolution for efficient therapeutics delivery. Drug Deliv. 2015;1-13.  [PubMed]  [DOI]  [Cited in This Article: ]
175.  Luppi B, Bigucci F, Corace G, Delucca A, Cerchiara T, Sorrenti M, Catenacci L, Di Pietra AM, Zecchi V. Albumin nanoparticles carrying cyclodextrins for nasal delivery of the anti-Alzheimer drug tacrine. Eur J Pharm Sci. 2011;44:559-565.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 92]  [Cited by in F6Publishing: 66]  [Article Influence: 5.1]  [Reference Citation Analysis (0)]
176.  Gao X, Chen J, Tao W, Zhu J, Zhang Q, Chen H, Jiang X. UEA I-bearing nanoparticles for brain delivery following intranasal administration. Int J Pharm. 2007;340:207-215.  [PubMed]  [DOI]  [Cited in This Article: ]
177.  Sood S, Jain K, Gowthamarajan K. Optimization of curcumin nanoemulsion for intranasal delivery using design of experiment and its toxicity assessment. Colloids Surf B Biointerfaces. 2014;113:330-337.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 176]  [Cited by in F6Publishing: 185]  [Article Influence: 16.8]  [Reference Citation Analysis (0)]
178.  Kameshima N, Nanjou T, Fukuhara T, Yanagisawa D, Tooyama I. Correlation of Aβ deposition in the nasal cavity with the formation of senile plaques in the brain of a transgenic mouse model of Alzheimer’s disease. Neurosci Lett. 2012;513:166-169.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 13]  [Cited by in F6Publishing: 15]  [Article Influence: 1.3]  [Reference Citation Analysis (0)]
179.  Kameshima N, Yanagisawa D, Tooyama I. β-Amyloid peptide (1-40) in the brain reaches the nasal cavity via a non-blood pathway. Neurosci Res. 2013;76:169-172.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 6]  [Cited by in F6Publishing: 7]  [Article Influence: 0.6]  [Reference Citation Analysis (0)]
180.  Hölscher C. First clinical data of the neuroprotective effects of nasal insulin application in patients with Alzheimer’s disease. Alzheimers Dement. 2014;10:S33-S37.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 50]  [Cited by in F6Publishing: 47]  [Article Influence: 4.7]  [Reference Citation Analysis (0)]
181.  Lee ST, Chu K, Jung KH, Kim JH, Huh JY, Yoon H, Park DK, Lim JY, Kim JM, Jeon D. miR-206 regulates brain-derived neurotrophic factor in Alzheimer disease model. Ann Neurol. 2012;72:269-277.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 208]  [Cited by in F6Publishing: 243]  [Article Influence: 20.3]  [Reference Citation Analysis (0)]
182.  Wang WX, Rajeev BW, Stromberg AJ, Ren N, Tang G, Huang Q, Rigoutsos I, Nelson PT. The expression of microRNA miR-107 decreases early in Alzheimer’s disease and may accelerate disease progression through regulation of beta-site amyloid precursor protein-cleaving enzyme 1. J Neurosci. 2008;28:1213-1223.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 560]  [Cited by in F6Publishing: 607]  [Article Influence: 37.9]  [Reference Citation Analysis (0)]
183.  Magen I, Hornstein E. Oligonucleotide-based therapy for neurodegenerative diseases. Brain Res. 2014;1584:116-128.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 30]  [Cited by in F6Publishing: 32]  [Article Influence: 3.2]  [Reference Citation Analysis (0)]
184.  Kanazawa T, Akiyama F, Kakizaki S, Takashima Y, Seta Y. Delivery of siRNA to the brain using a combination of nose-to-brain delivery and cell-penetrating peptide-modified nano-micelles. Biomaterials. 2013;34:9220-9226.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 150]  [Cited by in F6Publishing: 142]  [Article Influence: 12.9]  [Reference Citation Analysis (0)]
185.  Zhuang X, Xiang X, Grizzle W, Sun D, Zhang S, Axtell RC, Ju S, Mu J, Zhang L, Steinman L. Treatment of brain inflammatory diseases by delivering exosome encapsulated anti-inflammatory drugs from the nasal region to the brain. Mol Ther. 2011;19:1769-1779.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 820]  [Cited by in F6Publishing: 942]  [Article Influence: 72.5]  [Reference Citation Analysis (0)]
186.  Jordão JF, Ayala-Grosso CA, Markham K, Huang Y, Chopra R, McLaurin J, Hynynen K, Aubert I. Antibodies targeted to the brain with image-guided focused ultrasound reduces amyloid-beta plaque load in the TgCRND8 mouse model of Alzheimer’s disease. PLoS One. 2010;5:e10549.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 281]  [Cited by in F6Publishing: 285]  [Article Influence: 20.4]  [Reference Citation Analysis (0)]
187.  Jordão JF, Thévenot E, Markham-Coultes K, Scarcelli T, Weng YQ, Xhima K, O’Reilly M, Huang Y, McLaurin J, Hynynen K. Amyloid-β plaque reduction, endogenous antibody delivery and glial activation by brain-targeted, transcranial focused ultrasound. Exp Neurol. 2013;248:16-29.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 187]  [Cited by in F6Publishing: 230]  [Article Influence: 20.9]  [Reference Citation Analysis (0)]
188.  Nance E, Timbie K, Miller GW, Song J, Louttit C, Klibanov AL, Shih TY, Swaminathan G, Tamargo RJ, Woodworth GF. Non-invasive delivery of stealth, brain-penetrating nanoparticles across the blood-brain barrier using MRI-guided focused ultrasound. J Control Release. 2014;189:123-132.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 172]  [Cited by in F6Publishing: 176]  [Article Influence: 17.6]  [Reference Citation Analysis (0)]
189.  Treat LH, McDannold N, Zhang Y, Vykhodtseva N, Hynynen K. Improved anti-tumor effect of liposomal doxorubicin after targeted blood-brain barrier disruption by MRI-guided focused ultrasound in rat glioma. Ultrasound Med Biol. 2012;38:1716-1725.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 200]  [Cited by in F6Publishing: 220]  [Article Influence: 18.3]  [Reference Citation Analysis (0)]
190.  de la Torre JC. Carotid artery ultrasound and echocardiography testing to lower the prevalence of Alzheimer’s disease. J Stroke Cerebrovasc Dis. 2009;18:319-328.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 22]  [Cited by in F6Publishing: 26]  [Article Influence: 1.7]  [Reference Citation Analysis (0)]
191.  Roher AE, Garami Z, Tyas SL, Maarouf CL, Kokjohn TA, Belohlavek M, Vedders LJ, Connor D, Sabbagh MN, Beach TG. Transcranial doppler ultrasound blood flow velocity and pulsatility index as systemic indicators for Alzheimer’s disease. Alzheimers Dement. 2011;7:445-455.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 81]  [Cited by in F6Publishing: 99]  [Article Influence: 7.6]  [Reference Citation Analysis (0)]
192.  Santin MD, Debeir T, Bridal SL, Rooney T, Dhenain M. Fast in vivo imaging of amyloid plaques using μ-MRI Gd-staining combined with ultrasound-induced blood-brain barrier opening. Neuroimage. 2013;79:288-294.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 25]  [Cited by in F6Publishing: 27]  [Article Influence: 2.5]  [Reference Citation Analysis (0)]
193.  Xie F, Boska MD, Lof J, Uberti MG, Tsutsui JM, Porter TR. Effects of transcranial ultrasound and intravenous microbubbles on blood brain barrier permeability in a large animal model. Ultrasound Med Biol. 2008;34:2028-2034.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 60]  [Cited by in F6Publishing: 47]  [Article Influence: 2.9]  [Reference Citation Analysis (0)]
194.  Gratieri T, Alberti I, Lapteva M, Kalia YN. Next generation intra- and transdermal therapeutic systems: using non- and minimally-invasive technologies to increase drug delivery into and across the skin. Eur J Pharm Sci. 2013;50:609-622.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 81]  [Cited by in F6Publishing: 70]  [Article Influence: 6.4]  [Reference Citation Analysis (0)]
195.  Bastiat G, Plourde F, Motulsky A, Furtos A, Dumont Y, Quirion R, Fuhrmann G, Leroux JC. Tyrosine-based rivastigmine-loaded organogels in the treatment of Alzheimer’s disease. Biomaterials. 2010;31:6031-6038.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 61]  [Cited by in F6Publishing: 64]  [Article Influence: 4.6]  [Reference Citation Analysis (0)]
196.  Matsuoka Y, Gray AJ, Hirata-Fukae C, Minami SS, Waterhouse EG, Mattson MP, LaFerla FM, Gozes I, Aisen PS. Intranasal NAP administration reduces accumulation of amyloid peptide and tau hyperphosphorylation in a transgenic mouse model of Alzheimer’s disease at early pathological stage. J Mol Neurosci. 2007;31:165-170.  [PubMed]  [DOI]  [Cited in This Article: ]
197.  Gozes I, Giladi E, Pinhasov A, Bardea A, Brenneman DE. Activity-dependent neurotrophic factor: intranasal administration of femtomolar-acting peptides improve performance in a water maze. J Pharmacol Exp Ther. 2000;293:1091-1098.  [PubMed]  [DOI]  [Cited in This Article: ]
198.  Rat D, Schmitt U, Tippmann F, Dewachter I, Theunis C, Wieczerzak E, Postina R, van Leuven F, Fahrenholz F, Kojro E. Neuropeptide pituitary adenylate cyclase-activating polypeptide (PACAP) slows down Alzheimer’s disease-like pathology in amyloid precursor protein-transgenic mice. FASEB J. 2011;25:3208-3218.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 88]  [Cited by in F6Publishing: 103]  [Article Influence: 7.9]  [Reference Citation Analysis (0)]
199.  Vellas B, Carrie I, Gillette-Guyonnet S, Touchon J, Dantoine T, Dartigues JF, Cuffi MN, Bordes S, Gasnier Y, Robert P. MAPT Study: A Multidomain Approach for Preventing Alzheimer’s Disease: Design and Baseline Data. JPAD. 2014;1:3-44.  [PubMed]  [DOI]  [Cited in This Article: ]
200.  Anand R, Gill KD, Mahdi AA. Therapeutics of Alzheimer’s disease: Past, present and future. Neuropharmacology. 2014;76 Pt A:27-50.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 470]  [Cited by in F6Publishing: 500]  [Article Influence: 45.5]  [Reference Citation Analysis (0)]
201.  Imtiaz B, Tolppanen AM, Kivipelto M, Soininen H. Future directions in Alzheimer’s disease from risk factors to prevention. Biochem Pharmacol. 2014;88:661-670.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 138]  [Cited by in F6Publishing: 145]  [Article Influence: 14.5]  [Reference Citation Analysis (0)]
202.  Zilka N, Zilkova M, Kazmerova Z, Sarissky M, Cigankova V, Novak M. Mesenchymal stem cells rescue the Alzheimer’s disease cell model from cell death induced by misfolded truncated tau. Neuroscience. 2011;193:330-337.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 50]  [Cited by in F6Publishing: 53]  [Article Influence: 4.1]  [Reference Citation Analysis (0)]
203.  Kim KS, Kim HS, Park JM, Kim HW, Park MK, Lee HS, Lim DS, Lee TH, Chopp M, Moon J. Long-term immunomodulatory effect of amniotic stem cells in an Alzheimer’s disease model. Neurobiol Aging. 2013;34:2408-2420.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 107]  [Cited by in F6Publishing: 111]  [Article Influence: 10.1]  [Reference Citation Analysis (0)]
204.  Cummings JL, Morstorf T, Zhong K. Alzheimer’s disease drug-development pipeline: few candidates, frequent failures. Alzheimers Res Ther. 2014;6:37.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1236]  [Cited by in F6Publishing: 1141]  [Article Influence: 114.1]  [Reference Citation Analysis (0)]
205.  U.S. Food and Drug Administration. [Accessed 2014 Oct 20].  Available from: http://www.fda.gov/.  [PubMed]  [DOI]  [Cited in This Article: ]
206.  Cummings JL. Therapeutics. [Accessed 2014 Oct 20].  Available from: http://www.alzforum.org/therapeutics.  [PubMed]  [DOI]  [Cited in This Article: ]