Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2014 Aug 21.
Published in final edited form as: Expert Rev Neurother. 2010 May;10(5):711–728. doi: 10.1586/ern.10.29

Novel targets for Alzheimer's disease treatment

Joshua D Grill 1,*, Jeffrey L Cummings 1
PMCID: PMC4140224  NIHMSID: NIHMS594059  PMID: 20420492

Summary

Alzheimer's disease (AD) is a progressive neurodegenerative disease for which no cure exists. There is substantial need for new therapies that offer improved symptomatic benefit and disease-slowing capabilities. In recent decades there has been substantial progress in understanding the molecular and cellular changes associated with AD pathology. This has resulted in identification of a large number of new drug targets. These targets include but are not limited to therapies that aim to prevent production of or remove the beta amyloid (Aβ) protein that accumulates in neuritic plaques; prevent the hyperphosphorylation and aggregation into paired helical filaments of the microtubule-associated protein tau; and aim to keep neurons alive and functioning normally in the face of these pathologic challenges. We provide a review of these targets for drug development.

Keywords: Alzheimer's disease, dementia, treatment, beta amyloid, tau, therapeutics, beta secretase, gamma secretase, immunotherapy

Introduction

Recent research findings have led to greater understanding of disease neurobiology in Alzheimer's disease (AD) and identification of unique targets for drug development. Current therapeutic options aim at transmitter targets secondary to AD pathology. They offer limited efficacy and do not slow disease progression. The next generation of drugs for AD will alter the underlying disease course and/or provide greater symptomatic benefit. Targets for these drugs were identified in the study of AD pathophysiology and include but are not limited to the molecular events that result in the production and accumulation of the amyloid beta (Aβ) protein in neuritic plaques and the hyperphosphorylation, condensation and aggregation of the microtubule-associated protein tau in neurofibrillary tangles (NFTs). We review many of these targets, including those for which clinical development is ongoing or imminent. We begin by discussing current and emerging symptomatic targets,(Table 1) then we review targets with disease-modifying potential (Table 2).

Table 1.

Possible Drue Targets for Symptomatic Therapies for AD

Acetylcholinesterase inhbition*
NMDA receptor modulation*
Nicotinic acetylcholine receptor activation
Gamma amino butyric acid receptor blockade
Serotonin receptor activation and blocade
Histamine H3 receptor blockade
Phosphodiesterase inhibition
Brain metabolism enhancement
*

Targets for which FDA approved medications exist

Table 2.

Possible Drue Targets for Disease-Modifying Therapies for AD

Aβ Production
    Alpha secretase enhancement
    Beta secretase inhibition
    Gamma secretase inhibition
    Gamma secretase modulation
Aβ Degradation
    Neprilysin activation
    Insulin degrading enzyme activation
Aβ Removal
    Vaccination
    Passive immunization
    Receptor-mediated removal from CNS
    Prevent entry from periphery
Preventing Aβ Toxicity
    Prevent aggregation by binding Aβ
    Prevent oligomerization through metal protein attenuation
Tau
    Prevent tau aggregation
    Prevent tau hyperphosphorylation
    Facilitate tau phosphatases
    Microtubule stabilization
Neuroprotection
    Ensure neuronal health with growth factor treatment or growth factor receptor activation
    Prevent cell death with anti-apopotic agents
    Ensure mitochondrial health
    Block inflammatory disease processes

Targets Symptomatic Targets

Currently available therapies

Five medications are currently approved by the US Food and Drug Administration (FDA) for the treatment of AD. These agents, while providing important and useful symptomatic benefit, have limited or no impact on the underlying biology of AD, and no or limited impact on disease progression. Targets for these therapies are neurotransmitter-based. The acetylcholinesterase inhibitors (AChEIs) prevent the degradation of progressively decreasing acetylcholine (ACh) levels. AChEIs are effective throughout the course of AD.[1-8] The other class of approved medications includes only memantine, which regulates excitatory glutamatergic function and improves cognition in moderate-to-severe AD.[9,10] Administered together, these classes of medication may have enhanced symptomatic and long-term effects.[11,12] Altered neurotransmission in AD provides an array of targets for new drugs to improve cognition in the face of pathological challenge.

Nicotinic receptor agonism

The cholinergic system is well studied in AD. Neuronal nicotinic receptors (NNRs) for ACh are decreased in the AD brain[13,14] and this reduction correlates with cognitive impairment.[15] Most NNR agonists cross the blood-brain barrier.[16,17] NNR agonists improve cognition in small uncontrolled studies of normal volunteers[18] and AD patients.[19] Animal model studies suggest that these agents are as effective in improving cognitive performance as AChEI therapy.[20] Larger, more well-controlled clinical trials of NNR agonists will inform of their safety and efficacy in AD. No agents have been approved with NNRs as their target. A recent completed trial of one agent, AB-108, failed to meet its cognitive endpoints.

GABA receptor agonism and antagonism

As neurons that supply excitatory neurotransmission are lost in AD, inhibitory systems such as gamma aminobutyric acid (GABA) neurotransmission may go unchecked and result in cognitive impairment.[21] There is disagreement in the literature whether GABAergic neurons are less susceptible to the pathology of AD,[22-24] relative to excitatory neurons, or if levels of the neurotransmitter [25] and both GABAA and GABAB receptors[26] are decreased in AD. Regardless, agents that stabilize this system may modulate excitatory tone and reduce excitotoxicity. One selective GABAB receptor antagonist was clinically investigated in mild cognitive impairment,[27] but subsequently failed to demonstrate efficacy in AD.[28] Alternatively, inverse agonists of the GABAA receptor benzodiazepine-binding site (believed to reduce synaptic inhibition by reducing chloride ion flux) may enhance cognition.[29] Off target effects such as seizures and anxiety are a concern in compounds with inverse agonist or antagonist properties at the GABA receptor. A Phase II clinical trial of etazolate, a GABAA receptor modulator, recently completed enrollment. The synthetic GABA analog 3-amino-1-propanesulfonic acid (3APS; also known as tramiprosate or Alzhemed), was actively pursued as a treatment for AD, based more on its affinity for Aβ than any GABAergic properties.[30]

Serotonergic modulation

Of the 14 different serotonin (5-HT) receptor subtypes, several—including 5-HT1A, 5-HT4, and 5-HT6—are expressed in areas important to learning and memory[31] and significantly decline in AD.[32-36] Levels of 5-HTRs may also change in very mild disease,[33,37,38] further implicating these receptors in the cognitive impairment associated with AD. 5-HTRs may also be altered in other dementias.[39] The role of the 5-HTRs is complex, including regulation of a variety of other neurotransmitter systems, such as ACh, dopamine, GABA, and glutamate.[31] In line with this complexity, both selective agonists [40,41] and antagonists [42-45] of 5-HTRs have beneficial cognitive effects in animal models. One selective 5-HT1AR antagonist has reached Phase I investigation in man as a treatment for AD.[46] Results from a Phase II trial of a 5-HT6 antagonist are encouraging.[47]

Histamine H3 receptor antagonism

Four subtypes of histamine receptor exist and among them the H3 receptor is highly expressed in human brain, including in areas important to memory function.[48] H3 functions as a presynaptic autoreceptor and a postsynaptic heteroreceptor and regulates histamine function as well as release of a variety of neurotransmitters. Antagonists and inverse agonists of the H3 receptor increase neurotransmitter levels of ACh, dopamine, norepinephrine, and 5-HT.[49] Expression of H3 receptors in the periphery is limited, relative to expression in the brain. Unlike many other receptor systems, H3 receptor levels are maintained through the course of AD progression.[50] Development of H3 antagonists is therefore a promising target for AD therapies as well as a variety of other CNS disorders.[51] Clinical trials of H3 antagonists in AD are now underway.[49](www.clinicaltrials.gov)

Phosphodiesterase inhibition

Improving synaptic function may provide a target for symptomatic therapy in AD. The cyclic-AMP response element–binding protein (CREB) is a transcription factor that is important to memory[52,53] and down-regulated in models of altered synaptic function related to AD.[54] In transgenic animal models of AD, synaptic long-term potentiation (LTP) can be restored through enhancement of the CREB pathway by increasing cGMP through inhibition of phosphodiesterase (PDE). This appears to be true for pan-PDE inhibitors such as caffeine,[55] and for more specific PDE inhibitors.[56,57] New inhibitors of specific PDE subtypes have been identified,[58] but previously approved agents such as those indicated for treatment of erectile dysfunction are also being investigated.[57,59] The wide use of the latter suggests the likely safety of this class of agents in elderly subjects.

Metabolic enhancement

Brain hypometabolism is an early consequence of AD pathology.[60,61] The early occurrence of cerebral hypometabolism in AD makes it unlikely to result from neuronal loss but instead a result of synaptic dysfunction. Therefore, increasing glucose metabolism in the AD brain may provide symptomatic improvement. Glucose administration improves memory performance in healthy elderly participants.[62] In AD patients, however, glucose administration failed to improve cognitive performance.[63] Provision of alternative energy supplies such as ketogenic substances may improve brain metabolism and cognition.[64] One ketogenic dietary agent has recently been shown in a double-blind placebo-controlled trial to improve symptoms in AD,[65] and has subsequently received marketing permission from the U.S. Food and Drug Administration as a medical food.

Treatment with insulin also improves memory performance in AD.[63] The dysregulation of insulin in diabetes appears to increase risk for AD[66] and the insulin-degrading enzyme (IDE; also known as insulysin) also degrades Aβ.[67,68] These facts led to the investigation of the peroxisome proliferator-activated receptor-γ (PPAR-γ) agonist rosiglitazone as a treatment for AD.[69] Larger trials of rosiglitazone were made more difficult by cardiovascular health risks for this class of agents and one trial recently failed to demonstrate efficacy.[70] Direct intranasal administration of insulin also can improve cognitive function in AD.[71] Larger studies will be needed to confirm the symptomatic efficacy of intranasal insulin administration.

Disease Modifying Targets

Beta Amyloid

With extensive basic and clinical science support, the amyloid hypothesis remains the most prominent theory of AD pathology and target for drug development.[72-76] Central to this hypothesis is the proteolytic processing of the amyloid precursor protein (APP).[77] Posttranslational modification of APP occurs through two paths, beginning with cleavage by either α- or β-site enzymes. In either case, the produced C-terminal fragment is membrane-tethered and further processed within the transmembrane region by gamma secretase. Gamma secretase cleavage of the α-secretase product yields an apparently non-toxic protein fragment (p3), whereas the product of sequential beta-site cleavage enzyme (BACE) and gamma secretase cleavages is the 38-43 amino acid fibrillogenic amyloid beta protein (Aβ).[77] The 42-amino acid Aβ42 is most prone to aggregation, found in the core of neuritic plaques, and—when aggregated into oligomers—is largely considered the key pathological component in AD. Deposits of Aβ can occur anywhere that APP, BACE, and gamma secretase are found, and this includes both intraand extracellular spaces.[78] Aggregates of Aβ are toxic to synapses and neurons. Initial stages of aggregation, ranging from dimeric to dodecomeric soluble aggregates, are believed to be most toxic and are collectively referred to as Aβ oligomers.[79]

The amyloid hypothesis is supported by the largely identical clinical and pathologic phenotype among the sporadic late onset AD and the rare inherited early onset form of AD that results from autosomal dominant mutations to the genes for APP and two of the subunits in gamma secretase (presinilin 1 and 2).[73,80-82] Over 150 mutations that result in this familial AD (fAD) have now been identified, all of which exclusively involve these three genes integrally involved in Aβ production.[81] The largest number of disease-modifying drug candidates for AD target APP processing and Aβ: limiting production of the toxic form of Aβ; removing Aβ from the brain, through degradation, transport, or immunotherapeutic mobilization; or preventing the toxic effects of Aβ.

Aβ Production

Alpha Secretase activation

Increasing production of non-amyloidogenic products of APP processing through activation of α-secretase could provide a drug target for reducing pathologic Aβ. Few compounds have been identified that successfully activate α-secretase. Among those that have, receptor agonists for subtypes of muscarinic ACh receptors (mAChR) are appealing. It is clear that among the five different mAChR subtypes, M1 is highly expressed in hippocampus and cortex and involved in cognition.[83] Loss of M1 function induces cognitive impairment, and agonism of the M1 receptor is a logical target for cognitive enhancement.[84] M1 (and M3) receptor agonism also regulates APP processing by increasing α-secretase activity and inhibiting α-secretase.[85,86] It is not yet known if mAChR function in AD is impaired and will limit the efficacy of selective agents that target M1.[87] Treatment of AD patients with an M1 agonist, however, reduces CSF Aβ[88] and M1 agonists are in development for AD.[87]

Beta Secretase inhibition

Cleavage of APP by BACE (also known as α-secretase and memapsin-2) is the first step in the proteolytic processing of APP, making it the ideal position in the cascade to intervene and halt production of all posttranslational products. BACE activity is increased in sporadic forms of AD.[89-91] BACE knockouts have limited phenotypic changes beyond reduced levels of Aβ,[92,93] suggesting that potent inhibitors may have limited side-effect profiles. Development of agents capable of inhibiting the multisite activity of BACE on its APP substrate has been difficult. Successful inhibition initially required large molecules (>500KDa) that were unable to cross the blood-brain barrier.[94] Highly lipophilic, smaller orally available agents that have access to the CNS have recently been developed, however, and such agents are currently in human clinical trials.[95]

Gamma Secretase inhibition

The rationale for targeting gamma secretase is similar to that of BACE; prevention of gamma secretase action inhibits posttranslational processing of APP and formation of Aβ. Gamma secretase is a four-part complex, consisting of the membrane proteins presenilin-1 or -2, nicastrin, Aph-1, and Pen-2.(Reviewed in [96]) Of these, presenilin serves as the catalytic subunit. Presenilin-mediated catabolism is critical not only to APP proteolysis but also to a variety of other transmembrane proteins including the Notch signaling receptor. Cleavage of the Notch transmembrane receptor by presenilin permits nuclear translocation of the Notch signaling protein. Presenilin knockouts are lethal and phenotypically similar to Notch knockouts, suggesting the critical role of presenilin cleavage in Notch signaling.[97] The role of Notch in development is well described and includes regulating proliferation, cell fate decisions, and cellular growth.[98-100] In adult physiology Notch signaling continues to play a major role, and potent inhibition of Notch cleavage and signaling by presenilin/gamma secretase inhibitors may result in intolerable or dangerous adverse events related to gastrointestinal, lymphatic, skin, and immune system toxicity.[101] Therefore, targeting gamma secretase is a binary challenge. Activity must sufficiently lower Aβ to have clinical impact (the degree of Aβ-lowering to accomplish this is not yet known). Moreover, this activity must be selective to brain gamma secretase and not result in intolerability or dangerous side effects due to Notch toxicity. To this point, the primary strategy to balance efficacy and tolerability has been to pursue agents with minimal Notch activity. Multiple agents have now entered clinical trials. [102-104] Gamma secretase inhibitors are capable of reducing plasma Aβ[105] and lowering CSF production of Aβ[106,107] in AD patients and normal volunteers, respectively. LY450139 is a gamma secretase inhibitor currently in Phase III clinical trials.(www.clinicaltrials.gov)

Alternatively, agents that modulate gamma secretase (without directly preventing cleavage) may selectively lower Aβ without altering Notch activity. Early studies identified a variety of nonsteroidal anti-inflammatory drugs (NSAIDS) and NSAID-like compounds that could reduce Aβ production in vitro, that did not act on Notch or on the cyclooxygenases.[108,109] Among these, R-flurbiprofen advanced through clinical development but failed to demonstrate efficacy in a large phase III clinical trial.[110,111] Brain penetration of the agent may have been inadequate. Although these results were disappointing, gamma secretase modulators may still hold promise as a therapeutic class. Target sites other than presenilin may be plausible and drugs that act through conformational inhibition, rather than direct catalytic inhibition, would not induce Notch-related toxicity.[112] Additionally, Kukar and colleagues recently demonstrated that some NSAIDs can modulate gamma secretase activity through APP binding, suggesting another potential path to reduce Aβ without Notch toxicity.[113] Drugs that block gamma secretase activity through substrate binding may have the added benefit of acting as Aβ aggregation inhibitors.

Degradation of Aβ

Once present in the brain, the toxic form of Aβ must either be degraded or removed to prevent the clinical development of dementia.[114] Multiple endogenous pathways for Aβ degradation exist and include neutral endopeptidase (also known as neprilysin)[115,116], IDE, endothelin-converting enzyme,[117,118] angiotensin-converting enzyme (ACE), and metalloproteinase 9. Of these, neprilysin and IDE are thought to be the primary regulators of Aβ degradation, as well as the optimal drug targets.[119] Levels of Aβ degrading enzymes are reduced in AD.[120] Further, transgenic animals that lack the proteases key to Aβ degradation show increased brain Aβ deposition in a gene dose-dependent fashion[121], whereas APP transgenic mice that overexpress the neprilysin transgene demonstrate increased Aβ degradation (seen as a reduction in total soluble Aβ and plaque burden).[122] These animals, however, showed no reduction of oligomeric Aβ and no improvement in memory performance relative to APP mice.[122] APP transgenic mice engineered to overexpress IDE and neprilysin have reduced Aβ levels, virtually no plaque formation, and reduced astrogliosis, microgliosis, and dystrophic neurites.[123] These transgenic animals also show improved spatial memory in at least some models.[124]

Increased levels of neprilysin through viral vector delivered gene expression can lower Aβ in mouse models of AD.[125,126] Alternatively, molecular regulators of neprilysin levels in vivo could provide a more easily accomplished pharmacologic target.[127] Cabrol and colleagues recently completed a high-throughput screen in which they identified multiple small molecule activators of IDE,[128] suggesting that pharmacological manipulation of Aβ degradation enzymes is a realistic target for disease modification in AD. Given the number of degrading enzymes, it is important to demonstrate that inhibiting one pathway is adequate for a therapeutic benefit. Alternatively, multiple degradative pathways could be targeted.

Removal of Aβ

Vaccination

Vaccination with the full length Aβ peptide decreases amyloid burden and abrogates learning and memory impairment in animal models of AD.[129,130] Recent animal model investigations suggest that vaccination can reduce amlyoid burden as well as neurofibrillary tangle pathology.[131] A clinical trial of Aβ vaccination (AN1792) in 300 mild AD patients was halted due to a 6% incidence of T-cell mediated meningoencephalitis.[132] Preliminary results suggested a clinical benefit in participants who received therapy and generated antibodies against Aβ.[133] Analysis of the primary outcomes at the time of trial interruption demonstrated no drug-placebo difference.[134] Long-term follow up of survivors suggested a benefit in activities of daily living, quantified with the Disability Assessment for Dementia scale, among antibody responders.[135] Long-term follow-up to autopsy of the first subjects to die, however, demonstrated that there was no impact on clinical progression to advanced dementia, despite removal of plaque burden in the cortex among antibody responders.[136,137] Further, initial pathological analysis suggested that while Aβ removal was successful, NFT pathology was unaltered.[138] These findings have sparked continued debate as to whether Aβ provides the appropriate target for AD therapies.[139] Nevertheless, clinical development of immunotherapies for AD, including active vaccinations, remains a major research focus. Full length Aβ vaccinations, as well as peptide fragment vaccinations, are in development.[28]

Passive immunization

Passive immunization attempts to provide the Aβ-lowering and potential clinical benefit of vaccination, while avoiding the T-cell mediated response. Transgenic mouse studies confirm that the beneficial effects of vaccination including removal of Aβ burden, reductions in aggregated tau, and amelioration of cognitive deficit can be achieved with passive antibody therapy.[140-142] The exact mechanism of action for antibody therapy remains elusive, though several not mutually exclusive hypotheses have basic science support. Activation of microglia enhances breakdown of Aβ. The occurrence of “moth-eaten” plaques and increased immunohistochemical evidence of microglial response in vaccinated antibody responders that have come to autopsy from the AN1792 study support the hypothesis that immunotherapy enhances microglial response against Aβ.[136] Similarly, microglial activation occurs in response to vaccination [143] and passive immunization[144] in transgenic models of AD and is critical to reducing Aβ burden.[145] Additionally, antibodies may trap Aβ in the brain blood vasculature and transport it to the periphery where it can be more effectively degraded. Deglycosylated antibodies, which fail to activate the microglial response, persist in their ability to remove Aβ from the brains of mouse models of AD.[146,147] Similarly, passive immunization of transgenic animals that lack the Fc receptor necessary for microglial activation exhibit Aβ removal.[148] A variety of antibody therapies are now in clinical development, with two monoclonal antibody therapies in Phase III trials—bapineuzumab and solanezumab.

Interestingly, plaques in the brains of untreated AD patients are decorated with IgG antibodies, and these antibodies appear to elicit a microglial response.[149] Although this is insufficient for staving off clinical impairment, an inverse relationship exists between IgG level and plaque burden.[149] Therapeutic use of naturally produced autoantibodies in the form of intravenous immunoglobulin (IVIg) as treatment for AD is a current area of clinical investigation. IVIg treatment increased plasma and decreased CSF levels of Aβ in a study of 8 AD patients.[150] IVIg is in a large multisite clinical trial in the US.

Aβ Transport

Beside age, the most well established risk factor for AD is the apolipoprotein E (ApoE) genetic status. Three isotypes of ApoE exists: ε2, ε3, and ε4. ε2 and ε4 are distinguished from ε3 by only single amino acid substitutions.[151] Persons carrying one or two copies of the ε4 allele are at increased risk to develop AD and to do so at an earlier age.[152] In those with AD[153] and those at risk for AD,[154] ε4 carrier status results in increased Aβ deposition in the brain. Apolipoproteins play a role in Aβ metabolism and transport.[155] Apolipoproteins do not cross the blood brain barrier (BBB),[151] but may regulate passage of Aβ between the CNS and the periphery. The extent to which the different isoforms of ApoE bind Aβ is debated.(reviewed in[155]) ApoE ε4 may increase passage of Aβ from blood to brain.[156] Mouse models that overproduce and deposit Aβ, when combined with transgenic animals that lack ApoE, have reduced Aβ deposition.[157] Synthesized compounds that prevent ApoE binding to Aβ reduce plaque deposition and memory impairment in animal models of AD.[158] This may provide an avenue for therapeutic intervention in AD; though benefit may depend on ε4 carrier status.

It is likely that ApoE carries Aβ to a receptor that facilitates passage across the BBB.[114] The low-density lipoprotein receptor-related protein (LRP) seems to be involved in such receptor-mediated passage of Aβ.[159] Antibodies against the LRP reduce Aβ efflux from the brain.[160] With age LRP expression is decreased and LRP-mediated Aβ efflux is reduced, increasing risk for Aβ build up and clinical onset of AD.[160] Mechanistically, LRP has been shown to bind, endocytose, and trancytose Aβ, though questions remain over LRP affinity for Aβ40 versus Aβ42.[161] Nevertheless, peripheral administration of soluble LRP in AD as a mechanism to pull Aβ out of the brain has been proposed as a potential therapy.[162]

Alternative to enhancing Aβ removal from the brain, it may be useful to prevent its entry to the CNS from peripheral sources. The receptor for advanced glycation end-products (RAGE) is a multiligand receptor expressed by a variety of cell types throughout the body[163] and brain.[164,165] RAGE binds Aβ with high affinity and expression of the receptor is increased in AD.[164,166] Preventing ligands from binding RAGE lowers brain levels of Aβ.[167] Therefore, a variety of compounds that aim to inhibit RAGE interaction with its endogenous ligands or provide soluble alternative receptor binding are in development as treatments for AD. One such compound is currently being investigated in a Phase II clinical trial.

Thus, brain levels of Aβ are determined by a complex equation including production, removal, and degradation. While production of Aβ is largely understood, removal and degradation are multifactorial the role of each possible pathway remains to be fully comprehended. The balance (or imbalance) between the mechanisms by which brain Aβ levels are maintained, and the causal link between this balance and AD remains uncertain and debated.[168]

Aβ Oligomerization

The most synaptotoxic form of Aβ is neither the constitutive monomeric form, nor the fibrillar form deposited in plaques.[79,169-171] The greatest Aβ-derived toxicity results from initial aggregation stages when monomeric Aβ first oligomerizes into dimers, trimers, and other high-molecular-weight combinations. Oligomer reduction is the most compelling target in the Aβ cascade; reducing fibrillar Aβ may have little or no cognitive benefit. Animals with reduced plaque burden but high oligomer levels have no improvement in cognitive function.[170,172] It is plausible that Aβ plaques act as an endogenous sink, removing toxic oligomeric Aβ from the parenchyma[170] or they may serve as reservoirs for oligomeric Aβ. Estimates of how much total Aβ is soluble at any one time are as low as 5%. In animal models, reduction of soluble Aβ levels by 25% reduced plaque formation but had no effect on plaque burden.[173]

The aggregation properties of Aβ are fairly well established[174-176] and provide anti-aggregation targets for drug development.[177](FIGURE) Glabe and colleagues recently identified unique inhibitors capable of preventing oligomerization, fibrillization, or both.[178] Further, single agents may be able to decrease oligomerization while promoting fibrillization.[179] These findings support the hypothesis that multiple pathways of Aβ aggregation exist and may need to be further explored to optimize drug development that aims to reduce oligomers.

Figure.

Figure

Pathological cascade of AD. (Aß – amyloid beta protein; APP – amyloid precursor protein; IDE – insulin degrading enzyme; LTP – long term potentiation; NFT – neurofibrillary tangle; p-tau – phosphorylated tau protein).

The only agent aiming to prevent Aβ aggregation thus far tested in a large Phase III trial—tramiprosate—failed to meet its primary endpoint. It remains unclear if this agent lacked efficacy because it prevented fibrillization but not oligomerization of Aβ, or if Aβ is an inappropriate target.[28] Additionally, trial irregularities—including abnormal variance among study sites and lower than expected placebo decline—make it difficult to draw conclusions about drug efficacy.

A variety of other agents that aim specifically to prevent the oligomerization of Aβ are also in development. Cyclohexanehexol (also known as AZD-103) prevents oligomerization in vitro,[180] abrogates Aβ oligomer–induced impairments to LTP,[181] and alleviated AD cognitive impairment in a mouse model.[181,182] AZD-103 was safely tolerated in a human Phase I trial.

Other compounds may similarly inhibit oligomerization of Aβ, including grape-derived polyphenols,[183,184] omega-3 fatty acids such as docesahaexonic acid (DHA),[185] and curcumin.[185] In addition to potentially being aggregation inhibitors, each of these compounds are potent antioxidants and anti-inflammatory agents.[186,187] Epidemiologic studies of dietary intake[188-191] and comparisons of geographic regions where consumption is high for these compounds[192,193] suggest that consistent consumption can reduce risk for AD. Supplementing the diets of transgenic animals with these compounds confirms that they can reduce AD pathological burden in the brain.[194-196] For these reasons each of these classes of compounds have been or will be tested in human clinical trials as therapies for AD.[80](www.clinicaltrials.gov)

Bush and colleagues[197] have proposed a metal-related hypothesis of Aβ aggregation. In this model, metal ions such as copper and zinc, released with synaptic activity, promote the oligomerization of monomeric Aβ.[198,199] Such metal ions are elevated in AD[200] and are localized to plaques.[201] Based on these findings, a metal protein attenuating compound (MPAC) was tested in a clinical trial.[197] PBT2 demonstrated initial efficacy in AD, lowering CSF levels of Aβ and mitigating cognitive deterioration relative to placebo.[202-204] The specific mechanism of action of PBT2 remains debated,[205] however, and roles for the drug beyond local metal chelation have been proposed. At least one report suggests that clioquinol (another MPAC) may act in the mitochondria to inhibit enzymatic activity believed to underlie cellular aging.[206]

Tau and Neurofibrillary Tangles

The second pathological hallmark of AD is the intracellular neurofibrillary tangle (NFT). NFTs are condensed cytoskeletal inclusions composed of hyperphosphorylated paired helical filaments of the microtubule-associated protein tau (MAPtau or tau). In contrast to the axon-specific expression of tau in developing and healthy neurons,[207], in AD hyperphosphorylated tau is translocated to the somatodendritic compartment. Phosphorylation of tau is critical to its function, but hyperphosphorylated tau no longer binds to microtubules, instead aggregating into paired helical filaments.[208] The result is a general instability of microtubules and disruption of axonal transport that leads to neuronal injury and cell death. Increased levels of phosphorylated or total tau in the CSF are strong indicators of neurodegenerative disease or injury.[209] Though it is argued that a stronger link exists between NFT topography and clinical phenotype,[210] development of therapies targeting NFTs has lagged behind those that target Aβ.[211] Therapies targeting tau aim to reduce, stabilize, or prevent hyperphosphorylation or aggregation of the protein.

Total Tau

Multiple basic science models suggest that reducing tau can alleviate Aβ-dependent or Aβ-independent cognitive impairment in neurodegenerative models.[212,213] Given that tau is a constitutive part of the cell, however, removing tau entirely is not likely to be a realistic target.

Tau Aggregation

Oligomers of tau are part of normal functioning for the microtubule-associated protein and departure from this oligomeric structure into more aggregated compounds may represent the pathological step in tau processing.[214] Classes of agents that may act to prevent tau aggregation include anthraquinones, polyphenols, aminothienopyridazine, and phenothiazines.[215] Wischik and colleagues have begun clinical development of the phenothiazine methylene blue as a treatment for AD.[216,217] The initial clinical trial of this agent failed to meet its prespecified endpoints but long-term observations and biomarker studies suggested possible benefit. Clinical development in AD continues. Methylene blue also decreases Aβ oligomers in vitro by increasing fibrillar but not monomeric Aβ.[179]

Tau Hyperphosphorylation

Most of the drug development aiming at tau has targeted reducing hyperphosphorylation. Phosphorylation of tau can occur at many unique sites of the protein and through multiple pathways.[218-220] Among the various kinases involved in hyperphosphorylation of tau, glycogen synthase kinase 3 beta (GSK-3β) and the cyclin-dependent kinase-5 (cdk5) both target multiple phosphorylation sites on tau,[221] suggesting therapeutic potential for preventing the formation of NFTs.[219] Interactions between GSK-3β and cdk5 exist, and will require further evaluation to optimize treatments aimed at these kinases.[222,223]

Lithium and valproate are commonly used agents that have inhibitory action on GSK-3β and may stabilize tau.[211] Small and open label studies have suggested efficacy of these agents as therapies for cognitive and behavioral symptoms in AD.[224-226] Larger more controlled studies have failed to confirm efficacy for both lithium[227] and valproate.[228]

Agents with more appealing tolerability and safety profiles but still able to inhibit GSK-3β and stabilize tau are likely to reach clinical investigation in the near future. Recent findings suggest that caffeine inhibits GSK-3β,[229] in addition to inhibiting PDE (see above). Given epidemiologic findings of decreased incidence of AD in heavy caffeine users[230] and efficacy of caffeine in Aβ transgenic animal models,[231,232] further exploration of caffeine as a therapy for AD is warranted.

Phosphatases dephosphorylate tau and may offer an alternative to inhibition of hyperphosphorylation. Protein phosphatase 2A (PP2A) has been described to play a large role in regulation of tau phosphorylation and is decreased in AD.[233-235] Inhibition of PP2A is sufficient to result in hyperphosphorylation of tau, formation of NFT-like structures, and memory impairment in animal models.[236,237] Multiple PP2As exist and agents that can increase activity of these phosphatases, perhaps by targeting the endogenous proteins that inhibit their activity, are logical drug candidates for AD treatment.[238,239]

NAP is an 8 amino acid peptide (with N-A-P representing the first three amino acids in the peptide) believed to represent the active component of the glial-derived activity dependent neuroprotective protein (ADNP).[240] Unlike other trophic factors, which work via receptor-based mechanisms of action, NAP enters the neuron and interacts directly with microtubules.[241] NAP has potent ability in vitro to rescue neurons from Aβ-induced cell death[242] and also can reduce tau phosphorylation.[243] NAP treatment has demonstrated efficacy in the triple transgenic mouse model of AD.[244,245] Intranasally administered NAP treatment can cross the BBB[246] and has reached clinical investigation.[240]

It remains unclear whether therapies that alleviate Aβ pathology only or tau pathology only will ameliorate (or prevent) the cognitive decline in AD.[212,247] It is also important to note that therapies that successfully target tau may have applications beyond AD. Tau hyperphosphorylation is common in other forms of dementia, creating the possibility that successful tau-stabilizing agents will offer therapeutic efficacy in a variety of neurodegenerative conditions, including the frontotemporal dementias, progressive supranuclear palsy, and corticobasal degeneration.[248,249]

Neuroprotection

Neuroprotective strategies aim to ensure cell health in the presence of disease-specific pathology. These include therapies that reduce inflammation or other downstream markers of AD pathobiology. For example, epidemiologic findings have suggested disease-preventing properties for the non-steroidal anti-inflammatory drugs (NSAIDs)[250,251] and the statins.[252,253] The mechanisms by which these agents reduce AD, if they do, are not entirely clear. The anti-inflammatory properties of NSAIDs and the cholesterol-lowering properties of statins both might be useful in preventing or treating AD. Both classes of agents, however, may also reduce Aβ.[254] Prospective randomized placebo-controlled studies of NSAIDs have thus far failed to demonstrate efficacy as treatments,[255-257] although some studies were halted for safety concerns and thus prevented full examination.[258] There is a more mixed literature relating to statins. The Religious Orders Study [259] and Cardiovascular Health Study[260] found no effect of statin use on AD occurrence or neuropathology, but Li and colleagues demonstrated a reduced risk for NFT burden in statin users when examining participants followed longitudinally.[261] A prospective trial of 67 mild-to-moderate AD patients randomized to atorvostatin or placebo suggested a clinical benefit with treatment.[262] As a result of these findings, a larger prospective trial of atorvastatin and donepezil was initiated but failed to show a drug-placebo difference.[263,264]

Mitochondrial dysfunction plays a clear role in cell death.[265,266] Cellular aging and neurodegenerative disease mutate mitochondrial DNA, alter mitochondrial membrane permeability, and generally impair mitochondrial function.[267,268] In AD, soluble Aβ enters the mitochondria,[269] and results in mitochondrial membrane dysfunction, Ca2+ entry, reduced energy production, and oxidative stress that may ultimately lead to neuronal cell death.[268,270] Agents that stabilize the mitochondrial membrane, prevent mitochondrial DNA mutation, remove reactive oxygen species, or improve mitochondrial function may prevent age-related cellular changes and counteract neurodegenerative disease, including AD. Dimebon (dimebolin, latrepirdine) is a non-selective antihistamine previously approved in Russia now being developed for AD and Huntington's disease that may target mitochondrial membranes.[271] In addition to low affinity effects on AChE and glutamate receptors, dimebon appears to stabilize mitochondrial permeability transition pores, preventing ionic influx.[267] Doody and colleagues recently published results from a clinical trial of this agent in 183 mild-to-moderate AD patients.[271] After 26 weeks patients receiving therapy showed improvement relative to placebo on the mini mental status examination (MMSE), Alzheimer's Disease Assessment Scale-cognitive subscale (ADAS-cog), Alzheimer's Disease Cooperative Study-Activities of Daily Living scale (ADCSADL), Clinician Interview-Based Impression of Change (CIBIC)-plus, and Neuropsychiatric Inventory (NPI) outcomes. Blinded extension study to 12 months[271] and open label extension to 18-months[272] demonstrated continued benefit of therapy. Phase III testing of this agent is on-going.

Histones are proteins that organize chromatin within the cell nucleus. Histone regulation occurs by a variety of epigenetic modifiers, including methylation, acetylation, phosphorylation, ubiquitination, and ribosylation.[273] Among these, the balance between histone acetyltransferase (HAT) and histone deacetylase (HDAC) activity may be particularly relevant to synaptic function and memory, and may become unbalanced in AD and other neurodegenerative diseases.[274] Moreover, reestablishing this balance or increasing histone tail acetylation through inhibition of HDACs may facilitate memory recovery through synapse formation and dendritic growth.[275] Multiple classes of HDACs exist, composed of at least 10 subtypes. Recent work suggests that the HDAC2, but not HDAC1, subfamily of deacetylases is critical to synaptic function and may be an appropriate target for development of new therapies.[276] One small trial of nicotinamide, a class III HDAC inhibitor, improved memory in a triple transgenic mouse model of AD.[277] A trial of this agent has been initiated in AD patients.(www.clinicaltrials.gov)

Neurotrophic factors have long been known to promote survival and growth of neurons during CNS development, leading to great interest in the therapeutic application of these factors in the diseased adult brain.[278] Despite in vitro and animal model successes, trophic factor therapies have thus far failed in clinical investigation, either as a result of lack of efficacy or intolerability, in amyotrophic lateral sclerosis, Parkinson's disease, Huntington's disease, and other neurodegenerative disorders. [278-280] Issues with growth factor tolerability likely result from potent but often nonspecific neuritogenic activity.[281,282] Recent development of more targeted delivery approaches will allow for further evaluation of neurotrophic factor efficacy. In AD, nerve growth factor (NGF) delivery to the trk-A receptor-expressing cholinergic neurons of nucleus basalis of Meynert (NBM), which are known to be susceptible early in disease,[283] may increase cell survival and enhance cholinergic supply to the neocortex and hippocampus.[284-286] An initial trial of intracerebroventricular infusion of NGF offered limited efficacy and notable side effects.[281] Implantation of autologous fibroblasts genetically engineered to express NGF reduced the rate of cognitive decline and increased regional cerebral metabolism.[284] Focal delivery of gene therapy using adeno-associated virus (AAV)-delivered neurotrophic factor DNA provides long-term expression of the transgene and protein production, with minimal trophic factor diffusion.[287-289] Clinical investigation of NGF gene therapy as a treatment for AD is on-going. Similarly, Cerebrolysin, a compound that mimics neurotrophic factor effects has been tested in a clinical trial and results suggest improved cognition in AD.[290]

Five- Year View

To understand the immediate future in AD drug development, assessment of the current status of the field is necessary. We are on the cusp of a tremendous health care crisis as a result of AD. Already the source of immense societal cost, the number of AD cases in the US and worldwide is expected to triple in the coming decades. We are also on the brink of drastic improvements in our armamentarium to treat and, perhaps, prevent AD. The field currently pursues new treatments as well as improvements in diagnosis. Simultaneous advances in these two arenas will allow us to at least delay the onset of this debilitating disorder through preclinical identification and initiation of disease-slowing treatments. Successful development of an agent capable of delaying AD onset by 5 years will reduce the total number of cases by half and an agent that can delay onset for 10 years will essentially eradicate AD.[291]

Given the urgent need for and strong focus on developing new classes of therapy for AD, there is opportunity for rapid translation of basic science findings into clinical investigation. Clinical trials in AD have met great challenges. A number of agents through to hold promise have reached large-scale trials of efficacy only to fail to meet primary outcomes. Some of these have been positive studies (able to demonstrate definitive outcome) of negative results. Other cases, however, are less clear whether the drug failed or the trial did. Questions have arisen about optimal design for trials of disease modifying therapies, placebo decline, statistical power, and study conduct. Moreover, every failed trial in AD causes researchers to ask “are we intervening early enough?” Many of these issues remain unresolved.

The next five years will almost certainly bring new findings that will enhance our target selection in AD. A large number of agents are now in the final phase of clinical development (Table 3) and in the next 5 years it is likely that we will be able to confirm efficacy for one or more agents capable of at least mildly slowing disease progression. When this occurs, clinical trials investigating how best to use these agents and when and in what order to initiate their use will be critical. Only after such therapies have completed clinical investigation, however, will we have new treatments to offer our patients. Thus, it is critical that all physicians who care for AD patients refer eligible patients for participation in clinical trials.

Table 3.

AD Drug Candidates Currently in Phase III Trials*

Drug Sponsor Mechanism of Action
Bapineuzumab Elan/Wyeth Passive immunotherapy
Solanezumab Eli Lilly Passive immunotherapy
LY-450139 Eli Lilly Gamma secretase inhibition
Dimebon Medivation Mitochondrial stabilization
Intravenous Immunoglobulin Baxter Passive immunotherapy

Finally, debate remains over the definitive cause of the cognitive impairment in AD. It is conceivable that only through large-scale studies of agents for which cellular and molecular mechanisms of action can be confirmed will such debates finally be put to rest. That is, if a drug efficiently and completely removes amyloid from the AD brain and this entirely halts disease progression, ameliorates cognitive decline, or reverses disease effects, it will suggest that amyloid, in and of itself, is the primary culprit in disease pathology leading to clinical phenomenology. While this example is an unlikely scenario, it is clear that we will also make great strides in better understanding the pathology of AD with the development of agents that successfully prevent Aβ production or oligomerization, remove Aβ from the brain entirely, prevent tau hyperphosphorylation or tau aggregation (and a great many other potential mechanisms).

Expert Commentary

AD is a relentlessly progressive neurodegenerative disease that is increasing in prevalence rapidly. Elucidation of the pathophysiology of AD has resulted in better understanding of the preclinical events that lead to ostensible dementia. This pathophysiology includes the molecular and cellular events that lead to and result from the formation of amyloid plaques and neurofibrillary tangles. A multitude of drug targets relate to AD biology have been identified, including those for which cognitive symptom improvement is the goal, and others for which disease slowing is the goal. Only through clinical testing of such agents will these targets be validated and true causal pathology of disease be confirmed.

KEY ISSUES – bulleted executive summary.

  • No effective disease-modifying therapy for AD exists

  • A multitude of targets for new AD therapies have been identified in recent years

  • Testing of new agents in clinical trials is the rate-limiting step to new therapies

  • Clinical examination of new agents may resolve some of the disputes over causal pathology in AD

REFERENCES

  • 1.Rogers SL, Doody RS, Mohs RC, Friedhoff LT. Arch Intern Med. 9. Vol. 158. Donepezil Study Group; 1998. Donepezil improves cognition and global function in Alzheimer disease: a 15-week, double-blind, placebo-controlled study. pp. 1021–1031. [DOI] [PubMed] [Google Scholar]
  • 2.Rogers SL, Doody RS, Pratt RD, Ieni JR. Long-term efficacy and safety of donepezil in the treatment of Alzheimer's disease: final analysis of a US multicentre open-label study. Eur Neuropsychopharmacol. 2000;10(3):195–203. doi: 10.1016/s0924-977x(00)00067-5. [DOI] [PubMed] [Google Scholar]
  • 3.Tariot PN, Cummings JL, Katz IR, et al. A randomized, double-blind, placebo-controlled study of the efficacy and safety of donepezil in patients with Alzheimer's disease in the nursing home setting. J Am Geriatr Soc. 2001;49(12):1590–1599. [PubMed] [Google Scholar]
  • 4.Tariot PN, Solomon PR, Morris JC, Kershaw P, Lilienfeld S, Ding C. Neurology. 12. Vol. 54. The Galantamine USA-10 Study Group; 2000. A 5-month, randomized, placebo-controlled trial of galantamine in AD. pp. 2269–2276. [DOI] [PubMed] [Google Scholar]
  • 5.Rosler M, Anand R, Cicin-Sain A, et al. Efficacy and safety of rivastigmine in patients with Alzheimer's disease: international randomised controlled trial. Bmj. 1999;318(7184):633–638. doi: 10.1136/bmj.318.7184.633. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Raskind MA, Peskind ER, Wessel T, Yuan W. Neurology. 12. Vol. 54. The Galantamine USA-1 Study Group; 2000. Galantamine in AD: A 6-month randomized, placebo-controlled trial with a 6-month extension. pp. 2261–2268. [DOI] [PubMed] [Google Scholar]
  • 7.Farlow M, Anand R, Messina J, Jr., Hartman R, Veach J. A 52-week study of the efficacy of rivastigmine in patients with mild to moderately severe Alzheimer's disease. Eur Neurol. 2000;44(4):236–241. doi: 10.1159/000008243. [DOI] [PubMed] [Google Scholar]
  • 8.Winblad B, Kilander L, Eriksson S, et al. Donepezil in patients with severe Alzheimer's disease: double-blind, parallel-group, placebo-controlled study. Lancet. 2006;367(9516):1057–1065. doi: 10.1016/S0140-6736(06)68350-5. [DOI] [PubMed] [Google Scholar]
  • 9.Tariot PN, Farlow MR, Grossberg GT, Graham SM, McDonald S, Gergel I. Memantine treatment in patients with moderate to severe Alzheimer disease already receiving donepezil: a randomized controlled trial. JAMA. 2004;291(3):317–324. doi: 10.1001/jama.291.3.317. [DOI] [PubMed] [Google Scholar]
  • 10.van Dyck CH, Tariot PN, Meyers B, Malca Resnick E. A 24-week randomized, controlled trial of memantine in patients with moderate-to-severe Alzheimer disease. Alzheimer Dis Assoc Disord. 2007;21(2):136–143. doi: 10.1097/WAD.0b013e318065c495. [DOI] [PubMed] [Google Scholar]
  • 11*.Lopez OL, Becker JT, Wahed AS, et al. Long-term effects of the concomitant use of memantine with cholinesterase inhibition in Alzheimer disease. J Neurol Neurosurg Psychiatry. 2009;80(6):600–607. doi: 10.1136/jnnp.2008.158964. [Study of over 900 subjects for an average duration greater than 5 years examining the long-term effects of approved AD therapies. Those receiving cholinesterase therapies had delayed time to placement in nursing homes and the effect was increased in those receiving concomittant therapy with both cholinesterase inhibitors and memantine.] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Atri A, Shaughnessy LW, Locascio JJ, Growdon JH. Long-term course and effectiveness of combination therapy in Alzheimer disease. Alzheimer Dis Assoc Disord. 2008;22(3):209–221. doi: 10.1097/WAD.0b013e31816653bc. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Nordberg A, Winblad B. Reduced number of [3H]nicotine and [3H]acetylcholine binding sites in the frontal cortex of Alzheimer brains. Neurosci Lett. 1986;72(1):115–119. doi: 10.1016/0304-3940(86)90629-4. [DOI] [PubMed] [Google Scholar]
  • 14.Sabbagh MN, Shah F, Reid RT, et al. Pathologic and nicotinic receptor binding differences between mild cognitive impairment, Alzheimer disease, and normal aging. Arch Neurol. 2006;63(12):1771–1776. doi: 10.1001/archneur.63.12.1771. [DOI] [PubMed] [Google Scholar]
  • 15.Kadir A, Almkvist O, Wall A, Langstrom B, Nordberg A. PET imaging of cortical 11C-nicotine binding correlates with the cognitive function of attention in Alzheimer's disease. Psychopharmacology (Berl) 2006;188(4):509–520. doi: 10.1007/s00213-006-0447-7. [DOI] [PubMed] [Google Scholar]
  • 16.Mazurov A, Hauser T, Miller CH. Selective alpha7 nicotinic acetylcholine receptor ligands. Curr Med Chem. 2006;13(13):1567–1584. doi: 10.2174/092986706777442011. [DOI] [PubMed] [Google Scholar]
  • 17.Haydar SN, Ghiron C, Bettinetti L, et al. SAR and biological evaluation of SEN12333/WAY-317538: Novel alpha 7 nicotinic acetylcholine receptor agonist. Bioorg Med Chem. 2009;17(14):5247–5258. doi: 10.1016/j.bmc.2009.05.040. [DOI] [PubMed] [Google Scholar]
  • 18.Dunbar GC, Kuchibhatla R. Cognitive enhancement in man with ispronicline, a nicotinic partial agonist. J Mol Neurosci. 2006;30(1-2):169–172. doi: 10.1385/JMN:30:1:169. [DOI] [PubMed] [Google Scholar]
  • 19.Potter A, Corwin J, Lang J, Piasecki M, Lenox R, Newhouse PA. Acute effects of the selective cholinergic channel activator (nicotinic agonist) ABT-418 in Alzheimer's disease. Psychopharmacology (Berl) 1999;142(4):334–342. doi: 10.1007/s002130050897. [DOI] [PubMed] [Google Scholar]
  • 20.Marighetto A, Valerio S, Desmedt A, Philippin JN, Trocme-Thibierge C, Morain P. Comparative effects of the alpha7 nicotinic partial agonist, S 24795, and the cholinesterase inhibitor, donepezil, against aging-related deficits in declarative and working memory in mice. Psychopharmacology (Berl) 2008;197(3):499–508. doi: 10.1007/s00213-007-1063-x. [DOI] [PubMed] [Google Scholar]
  • 21.Rissman RA, De Blas AL, Armstrong DM. GABA(A) receptors in aging and Alzheimer's disease. J Neurochem. 2007;103(4):1285–1292. doi: 10.1111/j.1471-4159.2007.04832.x. [DOI] [PubMed] [Google Scholar]
  • 22.Rossor MN, Garrett NJ, Johnson AL, Mountjoy CQ, Roth M, Iversen LL. A post-mortem study of the cholinergic and GABA systems in senile dementia. Brain. 1982;105(Pt 2):313–330. doi: 10.1093/brain/105.2.313. [DOI] [PubMed] [Google Scholar]
  • 23.Mountjoy CQ, Rossor MN, Iversen LL, Roth M. Correlation of cortical cholinergic and GABA deficits with quantitative neuropathological findings in senile dementia. Brain. 1984;107(Pt 2):507–518. doi: 10.1093/brain/107.2.507. [DOI] [PubMed] [Google Scholar]
  • 24.Lowe SL, Francis PT, Procter AW, Palmer AM, Davison AN, Bowen DM. Gamma-aminobutyric acid concentration in brain tissue at two stages of Alzheimer's disease. Brain. 1988;111(Pt 4):785–799. doi: 10.1093/brain/111.4.785. [DOI] [PubMed] [Google Scholar]
  • 25.Ellison DW, Beal MF, Mazurek MF, Bird ED, Martin JB. A postmortem study of amino acid neurotransmitters in Alzheimer's disease. Ann Neurol. 1986;20(5):616–621. doi: 10.1002/ana.410200510. [DOI] [PubMed] [Google Scholar]
  • 26.Chu DC, Penney JB, Jr., Young AB. Cortical GABAB and GABAA receptors in Alzheimer's disease: a quantitative autoradiographic study. Neurology. 1987;37(9):1454–1459. doi: 10.1212/wnl.37.9.1454. [DOI] [PubMed] [Google Scholar]
  • 27.Froestl W, Gallagher M, Jenkins H, et al. SGS742: the first GABA(B) receptor antagonist in clinical trials. Biochem Pharmacol. 2004;68(8):1479–1487. doi: 10.1016/j.bcp.2004.07.030. [DOI] [PubMed] [Google Scholar]
  • 28.Sabbagh MN. Drug development for Alzheimer's disease: Where are we now and where are we headed? Am J Geriatr Pharmacother. 2009;7(3):167–185. doi: 10.1016/j.amjopharm.2009.06.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Sternfeld F, Carling RW, Jelley RA, et al. Selective, orally active gamma-aminobutyric acidA alpha5 receptor inverse agonists as cognition enhancers. J Med Chem. 2004;47(9):2176–2179. doi: 10.1021/jm031076j. [DOI] [PubMed] [Google Scholar]
  • 30.Aisen PS, Saumier D, Briand R, et al. A Phase II study targeting amyloid-beta with 3APS in mild-to-moderate Alzheimer disease. Neurology. 2006;67(10):1757–1763. doi: 10.1212/01.wnl.0000244346.08950.64. [DOI] [PubMed] [Google Scholar]
  • 31.King MV, Marsden CA, Fone KC. A role for the 5-HT(1A), 5-HT(4) and 5-HT(6) receptors in learning and memory. Trends Pharmacol Sci. 2008 doi: 10.1016/j.tips.2008.07.001. [DOI] [PubMed] [Google Scholar]
  • 32.Kepe V, Barrio JR, Huang SC, et al. Serotonin 1A receptors in the living brain of Alzheimer's disease patients. Proc Natl Acad Sci U S A. 2006;103(3):702–707. doi: 10.1073/pnas.0510237103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Truchot L, Costes SN, Zimmer L, et al. Up-regulation of hippocampal serotonin metabolism in mild cognitive impairment. Neurology. 2007;69(10):1012–1017. doi: 10.1212/01.wnl.0000271377.52421.4a. [DOI] [PubMed] [Google Scholar]
  • 34.Reynolds GP, Mason SL, Meldrum A, et al. 5-Hydroxytryptamine (5-HT)4 receptors in post mortem human brain tissue: distribution, pharmacology and effects of neurodegenerative diseases. Br J Pharmacol. 1995;114(5):993–998. doi: 10.1111/j.1476-5381.1995.tb13303.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Lorke DE, Lu G, Cho E, Yew DT. Serotonin 5-HT2A and 5-HT6 receptors in the prefrontal cortex of Alzheimer and normal aging patients. BMC Neurosci. 2006;7:36. doi: 10.1186/1471-2202-7-36. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Lai MK, Tsang SW, Alder JT, et al. Loss of serotonin 5-HT2A receptors in the postmortem temporal cortex correlates with rate of cognitive decline in Alzheimer's disease. Psychopharmacology (Berl) 2005;179(3):673–677. doi: 10.1007/s00213-004-2077-2. [DOI] [PubMed] [Google Scholar]
  • 37.Verdurand M, Berod A, Le Bars D, Zimmer L. Effects of amyloid-beta peptides on the serotoninergic 5-HT(1A) receptors in the rat hippocampus. Neurobiol Aging. 2009 doi: 10.1016/j.neurobiolaging.2009.01.008. [DOI] [PubMed] [Google Scholar]
  • 38.Hasselbalch SG, Madsen K, Svarer C, et al. Reduced 5-HT2A receptor binding in patients with mild cognitive impairment. Neurobiol Aging. 2008;29(12):1830–1838. doi: 10.1016/j.neurobiolaging.2007.04.011. [DOI] [PubMed] [Google Scholar]
  • 39.Elliott MS, Ballard CG, Kalaria RN, Perry R, Hortobagyi T, Francis PT. Increased binding to 5-HT1A and 5-HT2A receptors is associated with large vessel infarction and relative preservation of cognition. Brain. 2009;132(Pt 7):1858–1865. doi: 10.1093/brain/awp069. [DOI] [PubMed] [Google Scholar]
  • 40.Carli M, Balducci C, Samanin R. Stimulation of 5-HT(1A) receptors in the dorsal raphe ameliorates the impairment of spatial learning caused by intrahippocampal 7-chloro kynurenic acid in naive and pretrained rats. Psychopharmacology (Berl) 2001;158(1):39–47. doi: 10.1007/s002130100837. [DOI] [PubMed] [Google Scholar]
  • 41.Fontana DJ, Daniels SE, Wong EH, Clark RD, Eglen RM. The effects of novel, selective 5-hydroxytryptamine (5-HT)4 receptor ligands in rat spatial navigation. Neuropharmacology. 1997;36(4-5):689–696. doi: 10.1016/s0028-3908(97)00055-5. [DOI] [PubMed] [Google Scholar]
  • 42.Harder JA, Maclean CJ, Alder JT, Francis PT, Ridley RM. The 5-HT1A antagonist, WAY 100635, ameliorates the cognitive impairment induced by fornix transection in the marmoset. Psychopharmacology (Berl) 1996;127(3):245–254. [PubMed] [Google Scholar]
  • 43.Schechter LE, Smith DL, Rosenzweig-Lipson S, et al. Lecozotan (SRA-333): a selective serotonin 1A receptor antagonist that enhances the stimulated release of glutamate and acetylcholine in the hippocampus and possesses cognitive-enhancing properties. J Pharmacol Exp Ther. 2005;314(3):1274–1289. doi: 10.1124/jpet.105.086363. [DOI] [PubMed] [Google Scholar]
  • 44.Foley AG, Murphy KJ, Hirst WD, et al. The 5-HT(6) receptor antagonist SB-271046 reverses scopolamine-disrupted consolidation of a passive avoidance task and ameliorates spatial task deficits in aged rats. Neuropsychopharmacology. 2004;29(1):93–100. doi: 10.1038/sj.npp.1300332. [DOI] [PubMed] [Google Scholar]
  • 45.Da Silva Costa V, Duchatelle P, Boulouard M, Dauphin F. Selective 5-HT6 receptor blockade improves spatial recognition memory and reverses age-related deficits in spatial recognition memory in the mouse. Neuropsychopharmacology. 2009;34(2):488–500. doi: 10.1038/npp.2008.94. [DOI] [PubMed] [Google Scholar]
  • 46.Patat A, Parks V, Raje S, Plotka A, Chassard D, Le Coz F. Safety, tolerability, pharmacokinetics and pharmacodynamics of ascending single and multiple doses of lecozotan in healthy young and elderly subjects. Br J Clin Pharmacol. 2009;67(3):299–308. doi: 10.1111/j.1365-2125.2008.03348.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Upton N, Chuang TT, Hunter AJ, Virley DJ. 5-HT6 receptor antagonists as novel cognitive enhancing agents for Alzheimer's disease. Neurotherapeutics. 2008;5(3):458–469. doi: 10.1016/j.nurt.2008.05.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Lovenberg TW, Roland BL, Wilson SJ, et al. Cloning and functional expression of the human histamine H3 receptor. Mol Pharmacol. 1999;55(6):1101–1107. [PubMed] [Google Scholar]
  • 49.Esbenshade TA, Browman KE, Bitner RS, Strakhova M, Cowart MD, Brioni JD. The histamine H3 receptor: an attractive target for the treatment of cognitive disorders. Br J Pharmacol. 2008;154(6):1166–1181. doi: 10.1038/bjp.2008.147. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Medhurst AD, Roberts JC, Lee J, et al. Characterization of histamine H3 receptors in Alzheimer's Disease brain and amyloid over-expressing TASTPM mice. Br J Pharmacol. 2009;157(1):130–138. doi: 10.1111/j.1476-5381.2008.00075.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Leurs R, Bakker RA, Timmerman H, de Esch IJ. The histamine H3 receptor: from gene cloning to H3 receptor drugs. Nat Rev Drug Discov. 2005;4(2):107–120. doi: 10.1038/nrd1631. [DOI] [PubMed] [Google Scholar]
  • 52.Tully T, Bourtchouladze R, Scott R, Tallman J. Targeting the CREB pathway for memory enhancers. Nat Rev Drug Discov. 2003;2(4):267–277. doi: 10.1038/nrd1061. [DOI] [PubMed] [Google Scholar]
  • 53.Barco A, Pittenger C, Kandel ER. CREB, memory enhancement and the treatment of memory disorders: promises, pitfalls and prospects. Expert Opin Ther Targets. 2003;7(1):101–114. doi: 10.1517/14728222.7.1.101. [DOI] [PubMed] [Google Scholar]
  • 54.Vitolo OV, Sant'Angelo A, Costanzo V, Battaglia F, Arancio O, Shelanski M. Amyloid beta -peptide inhibition of the PKA/CREB pathway and long-term potentiation: reversibility by drugs that enhance cAMP signaling. Proc Natl Acad Sci U S A. 2002;99(20):13217–13221. doi: 10.1073/pnas.172504199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Dall'Igna OP, Fett P, Gomes MW, Souza DO, Cunha RA, Lara DR. Caffeine and adenosine A(2a) receptor antagonists prevent beta-amyloid (25-35)-induced cognitive deficits in mice. Exp Neurol. 2007;203(1):241–245. doi: 10.1016/j.expneurol.2006.08.008. [DOI] [PubMed] [Google Scholar]
  • 56.Gong B, Vitolo OV, Trinchese F, Liu S, Shelanski M, Arancio O. Persistent improvement in synaptic and cognitive functions in an Alzheimer mouse model after rolipram treatment. J Clin Invest. 2004;114(11):1624–1634. doi: 10.1172/JCI22831. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Puzzo D, Staniszewski A, Deng SX, et al. Phosphodiesterase 5 inhibition improves synaptic function, memory, and amyloid-beta load in an Alzheimer's disease mouse model. J Neurosci. 2009;29(25):8075–8086. doi: 10.1523/JNEUROSCI.0864-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Xia M, Huang R, Guo V, et al. Identification of compounds that potentiate CREB signaling as possible enhancers of long-term memory. Proc Natl Acad Sci U S A. 2009;106(7):2412–2417. doi: 10.1073/pnas.0813020106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Schultheiss D, Muller SV, Nager W, et al. Central effects of sildenafil (Viagra) on auditory selective attention and verbal recognition memory in humans: a study with event-related brain potentials. World J Urol. 2001;19(1):46–50. doi: 10.1007/pl00007092. [DOI] [PubMed] [Google Scholar]
  • 60.Langbaum JB, Chen K, Lee W, et al. Categorical and correlational analyses of baseline fluorodeoxyglucose positron emission tomography images from the Alzheimer's Disease Neuroimaging Initiative (ADNI). Neuroimage. 2009;45(4):1107–1116. doi: 10.1016/j.neuroimage.2008.12.072. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Reiman EM, Caselli RJ, Yun LS, et al. Preclinical evidence of Alzheimer's disease in persons homozygous for the epsilon 4 allele for apolipoprotein E. N Engl J Med. 1996;334(12):752–758. doi: 10.1056/NEJM199603213341202. [DOI] [PubMed] [Google Scholar]
  • 62.Manning CA, Stone WS, Korol DL, Gold PE. Glucose enhancement of 24-h memory retrieval in healthy elderly humans. Behav Brain Res. 1998;93(1-2):71–76. doi: 10.1016/s0166-4328(97)00136-8. [DOI] [PubMed] [Google Scholar]
  • 63.Craft S, Asthana S, Newcomer JW, et al. Enhancement of memory in Alzheimer disease with insulin and somatostatin, but not glucose. Arch Gen Psychiatry. 1999;56(12):1135–1140. doi: 10.1001/archpsyc.56.12.1135. [DOI] [PubMed] [Google Scholar]
  • 64.Costantini LC, Barr LJ, Vogel JL, Henderson ST. Hypometabolism as a therapeutic target in Alzheimer's disease. BMC Neurosci. 2008;9(Suppl 2):S16. doi: 10.1186/1471-2202-9-S2-S16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Henderson ST, Vogel JL, Barr LJ, Garvin F, Jones JJ, Costantini LC. Study of the ketogenic agent AC-1202 in mild to moderate Alzheimer's disease: a randomized, double-blind, placebo-controlled, multicenter trial. Nutr Metab (Lond) 2009;6:31. doi: 10.1186/1743-7075-6-31. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Arvanitakis Z, Wilson RS, Bienias JL, Evans DA, Bennett DA. Diabetes mellitus and risk of Alzheimer disease and decline in cognitive function. Arch Neurol. 2004;61(5):661–666. doi: 10.1001/archneur.61.5.661. [DOI] [PubMed] [Google Scholar]
  • 67.Qiu WQ, Walsh DM, Ye Z, et al. Insulin-degrading enzyme regulates extracellular levels of amyloid beta-protein by degradation. J Biol Chem. 1998;273(49):32730–32738. doi: 10.1074/jbc.273.49.32730. [DOI] [PubMed] [Google Scholar]
  • 68.Vekrellis K, Ye Z, Qiu WQ, et al. Neurons regulate extracellular levels of amyloid beta-protein via proteolysis by insulin-degrading enzyme. J Neurosci. 2000;20(5):1657–1665. doi: 10.1523/JNEUROSCI.20-05-01657.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Risner ME, Saunders AM, Altman JF, et al. Efficacy of rosiglitazone in a genetically defined population with mild-to-moderate Alzheimer's disease. Pharmacogenomics J. 2006;6(4):246–254. doi: 10.1038/sj.tpj.6500369. [DOI] [PubMed] [Google Scholar]
  • 70.Gold M AC, Zvartau-Hind M, Ritchie S, Saudners S, Craft S, Landreth G, Linnamagi U, Sawchak S. Effects of Rosiglitazone as a Monotherapy in APOE4-stratified Subjects with Mild to Moderate Alzheimer’s Disease. ICAD. 2009 Poster O1-04-06. [Google Scholar]
  • 71.Reger MA, Watson GS, Green PS, et al. Intranasal insulin improves cognition and modulates beta-amyloid in early AD. Neurology. 2008;70(6):440–448. doi: 10.1212/01.WNL.0000265401.62434.36. [DOI] [PubMed] [Google Scholar]
  • 72.Selkoe DJ. The molecular pathology of Alzheimer's disease. Neuron. 1991;6(4):487–498. doi: 10.1016/0896-6273(91)90052-2. [DOI] [PubMed] [Google Scholar]
  • 73.Hardy J. The amyloid hypothesis for Alzheimer's disease: a critical reappraisal. J Neurochem. 2009 doi: 10.1111/j.1471-4159.2009.06181.x. [DOI] [PubMed] [Google Scholar]
  • 74.Hardy J, Selkoe DJ. The amyloid hypothesis of Alzheimer's disease: progress and problems on the road to therapeutics. Science. 2002;297(5580):353–356. doi: 10.1126/science.1072994. [DOI] [PubMed] [Google Scholar]
  • 75.Hardy JA, Higgins GA. Alzheimer's disease: the amyloid cascade hypothesis. Science. 1992;256(5054):184–185. doi: 10.1126/science.1566067. [DOI] [PubMed] [Google Scholar]
  • 76.Glenner GG, Wong CW. Alzheimer's disease: initial report of the purification and characterization of a novel cerebrovascular amyloid protein. Biochem Biophys Res Commun. 1984;120(3):885–890. doi: 10.1016/s0006-291x(84)80190-4. [DOI] [PubMed] [Google Scholar]
  • 77.Thinakaran G, Koo EH. Amyloid precursor protein trafficking, processing, and function. J Biol Chem. 2008;283(44):29615–29619. doi: 10.1074/jbc.R800019200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.LaFerla FM, Green KN, Oddo S. Intracellular amyloid-beta in Alzheimer's disease. Nat Rev Neurosci. 2007;8(7):499–509. doi: 10.1038/nrn2168. [DOI] [PubMed] [Google Scholar]
  • 79.Shankar GM, Li S, Mehta TH, et al. Amyloid-beta protein dimers isolated directly from Alzheimer's brains impair synaptic plasticity and memory. Nat Med. 2008;14(8):837–842. doi: 10.1038/nm1782. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Ringman JM. What the study of persons at risk for familial Alzheimer's disease can tell us about the earliest stages of the disorder: a review. J Geriatr Psychiatry Neurol. 2005;18(4):228–233. doi: 10.1177/0891988705281878. [DOI] [PubMed] [Google Scholar]
  • 81.Rogaeva E, Kawarai T, George-Hyslop PS. Genetic complexity of Alzheimer's disease: successes and challenges. J Alzheimers Dis. 2006;9(3 Suppl):381–387. doi: 10.3233/jad-2006-9s343. [DOI] [PubMed] [Google Scholar]
  • 82.Blennow K, de Leon MJ, Zetterberg H. Alzheimer's disease. Lancet. 2006;368(9533):387–403. doi: 10.1016/S0140-6736(06)69113-7. [DOI] [PubMed] [Google Scholar]
  • 83.Levey AI, Kitt CA, Simonds WF, Price DL, Brann MR. Identification and localization of muscarinic acetylcholine receptor proteins in brain with subtype-specific antibodies. J Neurosci. 1991;11(10):3218–3226. doi: 10.1523/JNEUROSCI.11-10-03218.1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Anagnostaras SG, Murphy GG, Hamilton SE, et al. Selective cognitive dysfunction in acetylcholine M1 muscarinic receptor mutant mice. Nat Neurosci. 2003;6(1):51–58. doi: 10.1038/nn992. [DOI] [PubMed] [Google Scholar]
  • 85.Fisher A, Pittel Z, Haring R, et al. M1 muscarinic agonists can modulate some of the hallmarks in Alzheimer's disease: implications in future therapy. J Mol Neurosci. 2003;20(3):349–356. doi: 10.1385/JMN:20:3:349. [DOI] [PubMed] [Google Scholar]
  • 86.Caccamo A, Oddo S, Billings LM, et al. M1 receptors play a central role in modulating AD-like pathology in transgenic mice. Neuron. 2006;49(5):671–682. doi: 10.1016/j.neuron.2006.01.020. [DOI] [PubMed] [Google Scholar]
  • 87.Langmead CJ, Watson J, Reavill C. Muscarinic acetylcholine receptors as CNS drug targets. Pharmacol Ther. 2008;117(2):232–243. doi: 10.1016/j.pharmthera.2007.09.009. [DOI] [PubMed] [Google Scholar]
  • 88.Nitsch RM, Deng M, Tennis M, Schoenfeld D, Growdon JH. The selective muscarinic M1 agonist AF102B decreases levels of total Abeta in cerebrospinal fluid of patients with Alzheimer's disease. Ann Neurol. 2000;48(6):913–918. [PubMed] [Google Scholar]
  • 89.Holsinger RM, McLean CA, Beyreuther K, Masters CL, Evin G. Increased expression of the amyloid precursor beta-secretase in Alzheimer's disease. Ann Neurol. 2002;51(6):783–786. doi: 10.1002/ana.10208. [DOI] [PubMed] [Google Scholar]
  • 90.Yang LB, Lindholm K, Yan R, et al. Elevated beta-secretase expression and enzymatic activity detected in sporadic Alzheimer disease. Nat Med. 2003;9(1):3–4. doi: 10.1038/nm0103-3. [DOI] [PubMed] [Google Scholar]
  • 91.Fukumoto H, Cheung BS, Hyman BT, Irizarry MC. Beta-secretase protein and activity are increased in the neocortex in Alzheimer disease. Arch Neurol. 2002;59(9):1381–1389. doi: 10.1001/archneur.59.9.1381. [DOI] [PubMed] [Google Scholar]
  • 92.Luo Y, Bolon B, Kahn S, et al. Mice deficient in BACE1, the Alzheimer's beta-secretase, have normal phenotype and abolished beta-amyloid generation. Nat Neurosci. 2001;4(3):231–232. doi: 10.1038/85059. [DOI] [PubMed] [Google Scholar]
  • 93.Roberds SL, Anderson J, Basi G, et al. BACE knockout mice are healthy despite lacking the primary beta-secretase activity in brain: implications for Alzheimer's disease therapeutics. Hum Mol Genet. 2001;10(12):1317–1324. doi: 10.1093/hmg/10.12.1317. [DOI] [PubMed] [Google Scholar]
  • 94.Ghosh AK, Kumaragurubaran N, Hong L, et al. Design, synthesis, and X-ray structure of potent memapsin 2 (beta-secretase) inhibitors with isophthalamide derivatives as the P2-P3-ligands. J Med Chem. 2007;50(10):2399–2407. doi: 10.1021/jm061338s. [DOI] [PubMed] [Google Scholar]
  • 95.Ghosh AK, Gemma S, Tang J. beta-Secretase as a therapeutic target for Alzheimer's disease. Neurotherapeutics. 2008;5(3):399–408. doi: 10.1016/j.nurt.2008.05.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Wolfe MS. Inhibition and modulation of gamma-secretase for Alzheimer's disease. Neurotherapeutics. 2008;5(3):391–398. doi: 10.1016/j.nurt.2008.05.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Wong PC, Zheng H, Chen H, et al. Presenilin 1 is required for Notch1 and DII1 expression in the paraxial mesoderm. Nature. 1997;387(6630):288–292. doi: 10.1038/387288a0. [DOI] [PubMed] [Google Scholar]
  • 98.Louvi A, Artavanis-Tsakonas S. Notch signalling in vertebrate neural development. Nat Rev Neurosci. 2006;7(2):93–102. doi: 10.1038/nrn1847. [DOI] [PubMed] [Google Scholar]
  • 99.Lai EC. Notch signaling: control of cell communication and cell fate. Development. 2004;131(5):965–973. doi: 10.1242/dev.01074. [DOI] [PubMed] [Google Scholar]
  • 100.Lathia JD, Mattson MP, Cheng A. Notch: from neural development to neurological disorders. J Neurochem. 2008;107(6):1471–1481. doi: 10.1111/j.1471-4159.2008.05715.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Golde TE, Kukar TL. Medicine. Avoiding unintended toxicity. Science. 2009;324(5927):603–604. doi: 10.1126/science.1174267. [DOI] [PubMed] [Google Scholar]
  • 102.Siemers ER, Dean RA, Friedrich S, et al. Safety, tolerability, and effects on plasma and cerebrospinal fluid amyloid-beta after inhibition of gamma-secretase. Clin Neuropharmacol. 2007;30(6):317–325. doi: 10.1097/WNF.0b013e31805b7660. [DOI] [PubMed] [Google Scholar]
  • 103.Siemers ER, Quinn JF, Kaye J, et al. Effects of a gamma-secretase inhibitor in a randomized study of patients with Alzheimer disease. Neurology. 2006;66(4):602–604. doi: 10.1212/01.WNL.0000198762.41312.E1. [DOI] [PubMed] [Google Scholar]
  • 104.Imbimbo BP. Therapeutic potential of gamma-secretase inhibitors and modulators. Curr Top Med Chem. 2008;8(1):54–61. doi: 10.2174/156802608783334015. [DOI] [PubMed] [Google Scholar]
  • 105.Fleisher AS, Raman R, Siemers ER, et al. Phase 2 safety trial targeting amyloid beta production with a gamma-secretase inhibitor in Alzheimer disease. Arch Neurol. 2008;65(8):1031–1038. doi: 10.1001/archneur.65.8.1031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106*.Bateman RJ, Siemers ER, Mawuenyega KG, et al. A gamma-secretase inhibitor decreases amyloid-beta production in the central nervous system. Ann Neurol. 2009 doi: 10.1002/ana.21623. [Proof of concept demonstration of drug effect for a gamma secretase inhibitor in nondemented volunteers. Pulse-chase experiments utilizing stable isotope infusion and indwelling catheters were used to assess levels of CSF production of Aβ over 24 hours in response to treatment.] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Albright C DR, Olson R, Jere M, Slemmon R, Lentz K, Wang J, Denton R, Pilcher G, Zacaek R, Macor J, Wong O, Gu H, Berman R, Tong G. BMS-708163, a potent and selective γ-secretase inhibitor, decreases CSF Aβ at safe and tolerable doses in animals and humans. International Conference on Alzheimer's Disease. 2008 Abstract HT-01-05. [Google Scholar]
  • 108.Weggen S, Eriksen JL, Das P, et al. A subset of NSAIDs lower amyloidogenic Abeta42 independently of cyclooxygenase activity. Nature. 2001;414(6860):212–216. doi: 10.1038/35102591. [DOI] [PubMed] [Google Scholar]
  • 109.Weggen S, Eriksen JL, Sagi SA, et al. Evidence that nonsteroidal anti-inflammatory drugs decrease amyloid beta 42 production by direct modulation of gamma-secretase activity. J Biol Chem. 2003;278(34):31831–31837. doi: 10.1074/jbc.M303592200. [DOI] [PubMed] [Google Scholar]
  • 110.Hendrix SB, Wilcock GK. What we have learned from the Myriad trials. J Nutr Health Aging. 2009;13(4):362–364. doi: 10.1007/s12603-009-0044-7. [DOI] [PubMed] [Google Scholar]
  • 111.Green RC, Schneider LS, Amato DA, et al. Effect of Tarenflurbil on Cognitive Decline and Activities of Daily Living in Patients With Mild Alzheimer Disease: A Randomized Controlled Trial. JAMA. 2009;302(23):2557–2564. doi: 10.1001/jama.2009.1866. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Clarke EE, Churcher I, Ellis S, et al. Intra- or intercomplex binding to the gamma-secretase enzyme. A model to differentiate inhibitor classes. J Biol Chem. 2006;281(42):31279–31289. doi: 10.1074/jbc.M605051200. [DOI] [PubMed] [Google Scholar]
  • 113.Kukar TL, Ladd TB, Bann MA, et al. Substrate-targeting gamma-secretase modulators. Nature. 2008;453(7197):925–929. doi: 10.1038/nature07055. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Tanzi RE, Moir RD, Wagner SL. Clearance of Alzheimer's Abeta peptide: the many roads to perdition. Neuron. 2004;43(5):605–608. doi: 10.1016/j.neuron.2004.08.024. [DOI] [PubMed] [Google Scholar]
  • 115.Turner AJ, Tanzawa K. Mammalian membrane metallopeptidases: NEP, ECE, KELL, and PEX. FASEB J. 1997;11(5):355–364. doi: 10.1096/fasebj.11.5.9141502. [DOI] [PubMed] [Google Scholar]
  • 116.Iwata N, Tsubuki S, Takaki Y, et al. Identification of the major Abeta1-42-degrading catabolic pathway in brain parenchyma: suppression leads to biochemical and pathological deposition. Nat Med. 2000;6(2):143–150. doi: 10.1038/72237. [DOI] [PubMed] [Google Scholar]
  • 117.Eckman EA, Adams SK, Troendle FJ, et al. Regulation of steady-state beta-amyloid levels in the brain by neprilysin and endothelin-converting enzyme but not angiotensin-converting enzyme. J Biol Chem. 2006;281(41):30471–30478. doi: 10.1074/jbc.M605827200. [DOI] [PubMed] [Google Scholar]
  • 118.Eckman EA, Reed DK, Eckman CB. Degradation of the Alzheimer's amyloid beta peptide by endothelin-converting enzyme. J Biol Chem. 2001;276(27):24540–24548. doi: 10.1074/jbc.M007579200. [DOI] [PubMed] [Google Scholar]
  • 119.Nalivaeva NN, Fisk LR, Belyaev ND, Turner AJ. Amyloid-degrading enzymes as therapeutic targets in Alzheimer's disease. Curr Alzheimer Res. 2008;5(2):212–224. doi: 10.2174/156720508783954785. [DOI] [PubMed] [Google Scholar]
  • 120.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(1):115–121. doi: 10.1016/s0006-8993(01)03008-6. [DOI] [PubMed] [Google Scholar]
  • 121.Farris W, Mansourian S, Chang Y, et al. Insulin-degrading enzyme regulates the levels of insulin, amyloid beta-protein, and the beta-amyloid precursor protein intracellular domain in vivo. Proc Natl Acad Sci U S A. 2003;100(7):4162–4167. doi: 10.1073/pnas.0230450100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Meilandt WJ, Cisse M, Ho K, et al. Neprilysin overexpression inhibits plaque formation but fails to reduce pathogenic Abeta oligomers and associated cognitive deficits in human amyloid precursor protein transgenic mice. J Neurosci. 2009;29(7):1977–1986. doi: 10.1523/JNEUROSCI.2984-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Leissring MA, Farris W, Chang AY, et al. Enhanced proteolysis of beta-amyloid in APP transgenic mice prevents plaque formation, secondary pathology, and premature death. Neuron. 2003;40(6):1087–1093. doi: 10.1016/s0896-6273(03)00787-6. [DOI] [PubMed] [Google Scholar]
  • 124.Poirier R, Wolfer DP, Welzl H, et al. Neuronal neprilysin overexpression is associated with attenuation of Abeta-related spatial memory deficit. Neurobiol Dis. 2006;24(3):475–483. doi: 10.1016/j.nbd.2006.08.003. [DOI] [PubMed] [Google Scholar]
  • 125.Iwata N, Mizukami H, Shirotani K, et al. Presynaptic localization of neprilysin contributes to efficient clearance of amyloid-beta peptide in mouse brain. J Neurosci. 2004;24(4):991–998. doi: 10.1523/JNEUROSCI.4792-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Marr RA, Rockenstein E, Mukherjee A, et al. Neprilysin gene transfer reduces human amyloid pathology in transgenic mice. J Neurosci. 2003;23(6):1992–1996. doi: 10.1523/JNEUROSCI.23-06-01992.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Saito T, Iwata N, Tsubuki S, et al. Somatostatin regulates brain amyloid beta peptide Abeta42 through modulation of proteolytic degradation. Nat Med. 2005;11(4):434–439. doi: 10.1038/nm1206. [DOI] [PubMed] [Google Scholar]
  • 128.Cabrol C, Huzarska MA, Dinolfo C, et al. Small-molecule activators of insulin-degrading enzyme discovered through high-throughput compound screening. PLoS One. 2009;4(4):e5274. doi: 10.1371/journal.pone.0005274. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Schenk D, Barbour R, Dunn W, et al. Immunization with amyloid-beta attenuates Alzheimer-disease-like pathology in the PDAPP mouse. Nature. 1999;400(6740):173–177. doi: 10.1038/22124. [DOI] [PubMed] [Google Scholar]
  • 130.Morgan D, Diamond DM, Gottschall PE, et al. A beta peptide vaccination prevents memory loss in an animal model of Alzheimer's disease. Nature. 2000;408(6815):982–985. doi: 10.1038/35050116. [DOI] [PubMed] [Google Scholar]
  • 131.Wilcock DM, Gharkholonarehe N, Van Nostrand WE, Davis J, Vitek MP, Colton CA. Amyloid reduction by amyloid-beta vaccination also reduces mouse tau pathology and protects from neuron loss in two mouse models of Alzheimer's disease. J Neurosci. 2009;29(25):7957–7965. doi: 10.1523/JNEUROSCI.1339-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Orgogozo JM, Gilman S, Dartigues JF, et al. Subacute meningoencephalitis in a subset of patients with AD after Abeta42 immunization. Neurology. 2003;61(1):46–54. doi: 10.1212/01.wnl.0000073623.84147.a8. [DOI] [PubMed] [Google Scholar]
  • 133.Hock C, Konietzko U, Streffer JR, et al. Antibodies against beta-amyloid slow cognitive decline in Alzheimer's disease. Neuron. 2003;38(4):547–554. doi: 10.1016/s0896-6273(03)00294-0. [DOI] [PubMed] [Google Scholar]
  • 134.Gilman S, Koller M, Black RS, et al. Clinical effects of Abeta immunization (AN1792) in patients with AD in an interrupted trial. Neurology. 2005;64(9):1553–1562. doi: 10.1212/01.WNL.0000159740.16984.3C. [DOI] [PubMed] [Google Scholar]
  • 135.Vellas B, Black R, Thal LJ, et al. Long-term follow-up of patients immunized with AN1792: reduced functional decline in antibody responders. Curr Alzheimer Res. 2009;6(2):144–151. doi: 10.2174/156720509787602852. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136*.Holmes C, Boche D, Wilkinson D, et al. Long-term effects of Abeta42 immunisation in Alzheimer's disease: follow-up of a randomised, placebo-controlled phase I trial. Lancet. 2008;372(9634):216–223. doi: 10.1016/S0140-6736(08)61075-2. [Autopsy study of patients from the original Phase I trial of the AN-1792 vaccine four to 64 months after vaccination. Despite clearance of Aβ pathology in antibody responders, no differences were observed in survival or time to onset of severe dementia.] [DOI] [PubMed] [Google Scholar]
  • 137.Masliah E, Hansen L, Adame A, et al. Abeta vaccination effects on plaque pathology in the absence of encephalitis in Alzheimer disease. Neurology. 2005;64(1):129–131. doi: 10.1212/01.WNL.0000148590.39911.DF. [DOI] [PubMed] [Google Scholar]
  • 138.Nicoll JA, Wilkinson D, Holmes C, Steart P, Markham H, Weller RO. Neuropathology of human Alzheimer disease after immunization with amyloid-beta peptide: a case report. Nat Med. 2003;9(4):448–452. doi: 10.1038/nm840. [DOI] [PubMed] [Google Scholar]
  • 139.St George-Hyslop PH, Morris JC. Will anti-amyloid therapies work for Alzheimer's disease? Lancet. 2008;372(9634):180–182. doi: 10.1016/S0140-6736(08)61047-8. [DOI] [PubMed] [Google Scholar]
  • 140.Oddo S, Billings L, Kesslak JP, Cribbs DH, LaFerla FM. Abeta immunotherapy leads to clearance of early, but not late, hyperphosphorylated tau aggregates via the proteasome. Neuron. 2004;43(3):321–332. doi: 10.1016/j.neuron.2004.07.003. [DOI] [PubMed] [Google Scholar]
  • 141.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(26):6213–6220. doi: 10.1523/JNEUROSCI.0664-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Bard F, Cannon C, Barbour R, et al. Peripherally administered antibodies against amyloid beta-peptide enter the central nervous system and reduce pathology in a mouse model of Alzheimer disease. Nat Med. 2000;6(8):916–919. doi: 10.1038/78682. [DOI] [PubMed] [Google Scholar]
  • 143.Koenigsknecht-Talboo J, Meyer-Luehmann M, Parsadanian M, et al. Rapid microglial response around amyloid pathology after systemic anti-Abeta antibody administration in PDAPP mice. J Neurosci. 2008;28(52):14156–14164. doi: 10.1523/JNEUROSCI.4147-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Wilcock DM, DiCarlo G, Henderson D, et al. Intracranially administered anti-Abeta antibodies reduce beta-amyloid deposition by mechanisms both independent of and associated with microglial activation. J Neurosci. 2003;23(9):3745–3751. doi: 10.1523/JNEUROSCI.23-09-03745.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Wilcock DM, Munireddy SK, Rosenthal A, Ugen KE, Gordon MN, Morgan D. Microglial activation facilitates Abeta plaque removal following intracranial anti-Abeta antibody administration. Neurobiol Dis. 2004;15(1):11–20. doi: 10.1016/j.nbd.2003.09.015. [DOI] [PubMed] [Google Scholar]
  • 146.Takata K, Hirata-Fukae C, Becker AG, et al. Deglycosylated anti-amyloid beta antibodies reduce microglial phagocytosis and cytokine production while retaining the capacity to induce amyloid beta sequestration. Eur J Neurosci. 2007;26(9):2458–2468. doi: 10.1111/j.1460-9568.2007.05852.x. [DOI] [PubMed] [Google Scholar]
  • 147.Wilcock DM, Rojiani A, Rosenthal A, et al. Passive amyloid immunotherapy clears amyloid and transiently activates microglia in a transgenic mouse model of amyloid deposition. J Neurosci. 2004;24(27):6144–6151. doi: 10.1523/JNEUROSCI.1090-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Das P, Howard V, Loosbrock N, Dickson D, Murphy MP, Golde TE. Amyloid-beta immunization effectively reduces amyloid deposition in FcRgamma−/− knock-out mice. J Neurosci. 2003;23(24):8532–8538. doi: 10.1523/JNEUROSCI.23-24-08532.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Kellner A, Matschke J, Bernreuther C, Moch H, Ferrer I, Glatzel M. Autoantibodies against beta-amyloid are common in Alzheimer's disease and help control plaque burden. Ann Neurol. 2009;65(1):24–31. doi: 10.1002/ana.21475. [DOI] [PubMed] [Google Scholar]
  • 150.Relkin NR, Szabo P, Adamiak B, et al. 18-Month study of intravenous immunoglobulin for treatment of mild Alzheimer disease. Neurobiol Aging. 2008 doi: 10.1016/j.neurobiolaging.2007.12.021. [DOI] [PubMed] [Google Scholar]
  • 151.Ladu MJ, Reardon C, Van Eldik L, et al. Lipoproteins in the central nervous system. Ann N Y Acad Sci. 2000;903:167–175. doi: 10.1111/j.1749-6632.2000.tb06365.x. [DOI] [PubMed] [Google Scholar]
  • 152.Corder EH, Saunders AM, Strittmatter WJ, et al. Gene dose of apolipoprotein E type 4 allele and the risk of Alzheimer's disease in late onset families. Science. 1993;261(5123):921–923. doi: 10.1126/science.8346443. [DOI] [PubMed] [Google Scholar]
  • 153.Schmechel DE, Saunders AM, Strittmatter WJ, et al. Increased amyloid beta-peptide deposition in cerebral cortex as a consequence of apolipoprotein E genotype in late-onset Alzheimer disease. Proc Natl Acad Sci U S A. 1993;90(20):9649–9653. doi: 10.1073/pnas.90.20.9649. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Reiman EM, Chen K, Liu X, et al. Fibrillar amyloid-beta burden in cognitively normal people at 3 levels of genetic risk for Alzheimer's disease. Proc Natl Acad Sci U S A. 2009;106(16):6820–6825. doi: 10.1073/pnas.0900345106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Fan J, Donkin J, Wellington C. Greasing the wheels of Abeta clearance in Alzheimer's Disease: The role of lipids and apolipoprotein E. Biofactors. 2009;35(3):239–248. doi: 10.1002/biof.37. [DOI] [PubMed] [Google Scholar]
  • 156.Martel CL, Mackic JB, Matsubara E, et al. Isoform-specific effects of apolipoproteins E2, E3, and E4 on cerebral capillary sequestration and blood-brain barrier transport of circulating Alzheimer's amyloid beta. J Neurochem. 1997;69(5):1995–2004. doi: 10.1046/j.1471-4159.1997.69051995.x. [DOI] [PubMed] [Google Scholar]
  • 157.Bales KR, Verina T, Dodel RC, et al. Lack of apolipoprotein E dramatically reduces amyloid beta-peptide deposition. Nat Genet. 1997;17(3):263–264. doi: 10.1038/ng1197-263. [DOI] [PubMed] [Google Scholar]
  • 158.Sadowski MJ, Pankiewicz J, Scholtzova H, et al. Blocking the apolipoprotein E/amyloid-beta interaction as a potential therapeutic approach for Alzheimer's disease. Proc Natl Acad Sci U S A. 2006;103(49):18787–18792. doi: 10.1073/pnas.0604011103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Zlokovic BV. Clearing amyloid through the blood-brain barrier. J Neurochem. 2004;89(4):807–811. doi: 10.1111/j.1471-4159.2004.02385.x. [DOI] [PubMed] [Google Scholar]
  • 160.Shibata M, Yamada S, Kumar SR, et al. Clearance of Alzheimer's amyloid-ss(1-40) peptide from brain by LDL receptor-related protein-1 at the blood-brain barrier. J Clin Invest. 2000;106(12):1489–1499. doi: 10.1172/JCI10498. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Deane R, Wu Z, Sagare A, et al. LRP/amyloid beta-peptide interaction mediates differential brain efflux of Abeta isoforms. Neuron. 2004;43(3):333–344. doi: 10.1016/j.neuron.2004.07.017. [DOI] [PubMed] [Google Scholar]
  • 162.Sagare A, Deane R, Bell RD, et al. Clearance of amyloid-beta by circulating lipoprotein receptors. Nat Med. 2007;13(9):1029–1031. doi: 10.1038/nm1635. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Brett J, Schmidt AM, Yan SD, et al. Survey of the distribution of a newly characterized receptor for advanced glycation end products in tissues. Am J Pathol. 1993;143(6):1699–1712. [PMC free article] [PubMed] [Google Scholar]
  • 164.Lue LF, Walker DG, Brachova L, et al. Involvement of microglial receptor for advanced glycation endproducts (RAGE) in Alzheimer's disease: identification of a cellular activation mechanism. Exp Neurol. 2001;171(1):29–45. doi: 10.1006/exnr.2001.7732. [DOI] [PubMed] [Google Scholar]
  • 165.Sasaki N, Toki S, Chowei H, et al. Immunohistochemical distribution of the receptor for advanced glycation end products in neurons and astrocytes in Alzheimer's disease. Brain Res. 2001;888(2):256–262. doi: 10.1016/s0006-8993(00)03075-4. [DOI] [PubMed] [Google Scholar]
  • 166.Yan SD, Chen X, Fu J, et al. RAGE and amyloid-beta peptide neurotoxicity in Alzheimer's disease. Nature. 1996;382(6593):685–691. doi: 10.1038/382685a0. [DOI] [PubMed] [Google Scholar]
  • 167.Deane R, Du Yan S, Submamaryan RK, et al. RAGE mediates amyloid-beta peptide transport across the blood-brain barrier and accumulation in brain. Nat Med. 2003;9(7):907–913. doi: 10.1038/nm890. [DOI] [PubMed] [Google Scholar]
  • 168.Zlokovic BV, Yamada S, Holtzman D, Ghiso J, Frangione B. Clearance of amyloid beta-peptide from brain: transport or metabolism? Nat Med. 2000;6(7):718. doi: 10.1038/77397. [DOI] [PubMed] [Google Scholar]
  • 169.Lesne S, Koh MT, Kotilinek L, et al. A specific amyloid-beta protein assembly in the brain impairs memory. Nature. 2006;440(7082):352–357. doi: 10.1038/nature04533. [DOI] [PubMed] [Google Scholar]
  • 170.Cheng IH, Scearce-Levie K, Legleiter J, et al. Accelerating amyloid-beta fibrillization reduces oligomer levels and functional deficits in Alzheimer disease mouse models. J Biol Chem. 2007;282(33):23818–23828. doi: 10.1074/jbc.M701078200. [DOI] [PubMed] [Google Scholar]
  • 171.Cleary JP, Walsh DM, Hofmeister JJ, et al. Natural oligomers of the amyloid-beta protein specifically disrupt cognitive function. Nat Neurosci. 2005;8(1):79–84. doi: 10.1038/nn1372. [DOI] [PubMed] [Google Scholar]
  • 172.Head E, Pop V, Vasilevko V, et al. A two-year study with fibrillar beta-amyloid (Abeta) immunization in aged canines: effects on cognitive function and brain Abeta. J Neurosci. 2008;28(14):3555–3566. doi: 10.1523/JNEUROSCI.0208-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Yan P, Bero AW, Cirrito JR, et al. Characterizing the appearance and growth of amyloid plaques in APP/PS1 mice. J Neurosci. 2009;29(34):10706–10714. doi: 10.1523/JNEUROSCI.2637-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174.Burdick D, Soreghan B, Kwon M, et al. Assembly and aggregation properties of synthetic Alzheimer's A4/beta amyloid peptide analogs. J Biol Chem. 1992;267(1):546–554. [PubMed] [Google Scholar]
  • 175.Podlisny MB, Ostaszewski BL, Squazzo SL, et al. Aggregation of secreted amyloid beta-protein into sodium dodecyl sulfate-stable oligomers in cell culture. J Biol Chem. 1995;270(16):9564–9570. doi: 10.1074/jbc.270.16.9564. [DOI] [PubMed] [Google Scholar]
  • 176.Harmeier A, Wozny C, Rost BR, et al. Role of amyloid-beta glycine 33 in oligomerization, toxicity, and neuronal plasticity. J Neurosci. 2009;29(23):7582–7590. doi: 10.1523/JNEUROSCI.1336-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Walsh DM, Townsend M, Podlisny MB, et al. Certain inhibitors of synthetic amyloid beta-peptide (Abeta) fibrillogenesis block oligomerization of natural Abeta and thereby rescue long-term potentiation. J Neurosci. 2005;25(10):2455–2462. doi: 10.1523/JNEUROSCI.4391-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 178**.Necula M, Kayed R, Milton S, Glabe CG. Small molecule inhibitors of aggregation indicate that amyloid beta oligomerization and fibrillization pathways are independent and distinct. J Biol Chem. 2007;282(14):10311–10324. doi: 10.1074/jbc.M608207200. [In vitro demonstration that the pathway from Aβ monomer to Aβ plaque may not be a linear progression that includes the stage of oligomerization. Molecular compounds capable of inhibiting oligomerization, fibrilization, or both were identified.] [DOI] [PubMed] [Google Scholar]
  • 179.Necula M, Breydo L, Milton S, et al. Methylene blue inhibits amyloid Abeta oligomerization by promoting fibrillization. Biochemistry. 2007;46(30):8850–8860. doi: 10.1021/bi700411k. [DOI] [PubMed] [Google Scholar]
  • 180.McLaurin J, Franklin T, Chakrabartty A, Fraser PE. Phosphatidylinositol and inositol involvement in Alzheimer amyloid-beta fibril growth and arrest. J Mol Biol. 1998;278(1):183–194. doi: 10.1006/jmbi.1998.1677. [DOI] [PubMed] [Google Scholar]
  • 181.Townsend M, Cleary JP, Mehta T, et al. Orally available compound prevents deficits in memory caused by the Alzheimer amyloid-beta oligomers. Ann Neurol. 2006;60(6):668–676. doi: 10.1002/ana.21051. [DOI] [PubMed] [Google Scholar]
  • 182.McLaurin J, Kierstead ME, Brown ME, et al. Cyclohexanehexol inhibitors of Abeta aggregation prevent and reverse Alzheimer phenotype in a mouse model. Nat Med. 2006;12(7):801–808. doi: 10.1038/nm1423. [DOI] [PubMed] [Google Scholar]
  • 183.Wang J, Ho L, Zhao W, et al. Grape-derived polyphenolics prevent Abeta oligomerization and attenuate cognitive deterioration in a mouse model of Alzheimer's disease. J Neurosci. 2008;28(25):6388–6392. doi: 10.1523/JNEUROSCI.0364-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184.Ono K, Condron MM, Ho L, et al. Effects of grape seed-derived polyphenols on amyloid beta-protein self-assembly and cytotoxicity. J Biol Chem. 2008;283(47):32176–32187. doi: 10.1074/jbc.M806154200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Ma QL, Yang F, Rosario ER, et al. Beta-amyloid oligomers induce phosphorylation of tau and inactivation of insulin receptor substrate via c-Jun N-terminal kinase signaling: suppression by omega-3 fatty acids and curcumin. J Neurosci. 2009;29(28):9078–9089. doi: 10.1523/JNEUROSCI.1071-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Cole GM, Lim GP, Yang F, et al. Prevention of Alzheimer's disease: Omega-3 fatty acid and phenolic anti-oxidant interventions. Neurobiol Aging. 2005;26(Suppl 1):133–136. doi: 10.1016/j.neurobiolaging.2005.09.005. [DOI] [PubMed] [Google Scholar]
  • 187.Vingtdeux V, Dreses-Werringloer U, Zhao H, Davies P, Marambaud P. Therapeutic potential of resveratrol in Alzheimer's disease. BMC Neurosci. 2008;9(Suppl 2):S6. doi: 10.1186/1471-2202-9-S2-S6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Barberger-Gateau P, Letenneur L, Deschamps V, Peres K, Dartigues JF, Renaud S. Fish, meat, and risk of dementia: cohort study. BMJ. 2002;325(7370):932–933. doi: 10.1136/bmj.325.7370.932. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 189.Morris MC, Evans DA, Bienias JL, et al. Consumption of fish and n-3 fatty acids and risk of incident Alzheimer disease. Arch Neurol. 2003;60(7):940–946. doi: 10.1001/archneur.60.7.940. [DOI] [PubMed] [Google Scholar]
  • 190.Luchsinger JA, Tang MX, Siddiqui M, Shea S, Mayeux R. Alcohol intake and risk of dementia. J Am Geriatr Soc. 2004;52(4):540–546. doi: 10.1111/j.1532-5415.2004.52159.x. [DOI] [PubMed] [Google Scholar]
  • 191.Schaefer EJ, Bongard V, Beiser AS, et al. Plasma phosphatidylcholine docosahexaenoic acid content and risk of dementia and Alzheimer disease: the Framingham Heart Study. Arch Neurol. 2006;63(11):1545–1550. doi: 10.1001/archneur.63.11.1545. [DOI] [PubMed] [Google Scholar]
  • 192.Chandra V, Ganguli M, Pandav R, Johnston J, Belle S, DeKosky ST. Prevalence of Alzheimer's disease and other dementias in rural India: the Indo-US study. Neurology. 1998;51(4):1000–1008. doi: 10.1212/wnl.51.4.1000. [DOI] [PubMed] [Google Scholar]
  • 193.Chandra V, Pandav R, Dodge HH, et al. Incidence of Alzheimer's disease in a rural community in India: the Indo-US study. Neurology. 2001;57(6):985–989. doi: 10.1212/wnl.57.6.985. [DOI] [PubMed] [Google Scholar]
  • 194.Lim GP, Calon F, Morihara T, et al. A diet enriched with the omega-3 fatty acid docosahexaenoic acid reduces amyloid burden in an aged Alzheimer mouse model. J Neurosci. 2005;25(12):3032–3040. doi: 10.1523/JNEUROSCI.4225-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.Green KN, Martinez-Coria H, Khashwji H, et al. Dietary docosahexaenoic acid and docosapentaenoic acid ameliorate amyloid-beta and tau pathology via a mechanism involving presenilin 1 levels. J Neurosci. 2007;27(16):4385–4395. doi: 10.1523/JNEUROSCI.0055-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 196.Wang J, Ho L, Zhao Z, et al. Moderate consumption of Cabernet Sauvignon attenuates Abeta neuropathology in a mouse model of Alzheimer's disease. FASEB J. 2006;20(13):2313–2320. doi: 10.1096/fj.06-6281com. [DOI] [PubMed] [Google Scholar]
  • 197.Bush AI, Tanzi RE. Therapeutics for Alzheimer's disease based on the metal hypothesis. Neurotherapeutics. 2008;5(3):421–432. doi: 10.1016/j.nurt.2008.05.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 198.Bush AI, Pettingell WH, Jr., Paradis MD, Tanzi RE. Modulation of A beta adhesiveness and secretase site cleavage by zinc. J Biol Chem. 1994;269(16):12152–12158. [PubMed] [Google Scholar]
  • 199.Bush AI, Pettingell WH, Multhaup G, et al. Rapid induction of Alzheimer A beta amyloid formation by zinc. Science. 1994;265(5177):1464–1467. doi: 10.1126/science.8073293. [DOI] [PubMed] [Google Scholar]
  • 200.Lovell MA, Robertson JD, Teesdale WJ, Campbell JL, Markesbery WR. Copper, iron and zinc in Alzheimer's disease senile plaques. J Neurol Sci. 1998;158(1):47–52. doi: 10.1016/s0022-510x(98)00092-6. [DOI] [PubMed] [Google Scholar]
  • 201.Suh SW, Jensen KB, Jensen MS, et al. Histochemically-reactive zinc in amyloid plaques, angiopathy, and degenerating neurons of Alzheimer's diseased brains. Brain Res. 2000;852(2):274–278. doi: 10.1016/s0006-8993(99)02096-x. [DOI] [PubMed] [Google Scholar]
  • 202.Ibach B, Haen E, Marienhagen J, Hajak G. Clioquinol treatment in familiar early onset of Alzheimer's disease: a case report. Pharmacopsychiatry. 2005;38(4):178–179. doi: 10.1055/s-2005-871241. [DOI] [PubMed] [Google Scholar]
  • 203.Lannfelt L, Blennow K, Zetterberg H, et al. Safety, efficacy, and biomarker findings of PBT2 in targeting Abeta as a modifying therapy for Alzheimer's disease: a phase IIa, double-blind, randomised, placebo-controlled trial. Lancet Neurol. 2008;7(9):779–786. doi: 10.1016/S1474-4422(08)70167-4. [DOI] [PubMed] [Google Scholar]
  • 204.Ritchie CW, Bush AI, Mackinnon A, et al. 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(12):1685–1691. doi: 10.1001/archneur.60.12.1685. [DOI] [PubMed] [Google Scholar]
  • 205.Cahoon L. The curious case of clioquinol. Nat Med. 2009;15(4):356–359. doi: 10.1038/nm0409-356. [DOI] [PubMed] [Google Scholar]
  • 206.Wang Y, Branicky R, Stepanyan Z, et al. The anti-neurodegeneration drug clioquinol inhibits the aging-associated protein CLK-1. J Biol Chem. 2009;284(1):314–323. doi: 10.1074/jbc.M807579200. [DOI] [PubMed] [Google Scholar]
  • 207.Bradke F, Dotti CG. Establishment of neuronal polarity: lessons from cultured hippocampal neurons. Curr Opin Neurobiol. 2000;10(5):574–581. doi: 10.1016/s0959-4388(00)00124-0. [DOI] [PubMed] [Google Scholar]
  • 208.Lee VM, Trojanowski JQ. The disordered neuronal cytoskeleton in Alzheimer's disease. Curr Opin Neurobiol. 1992;2(5):653–656. doi: 10.1016/0959-4388(92)90034-i. [DOI] [PubMed] [Google Scholar]
  • 209.Clark CM, Xie S, Chittams J, et al. Cerebrospinal fluid tau and beta-amyloid: how well do these biomarkers reflect autopsy-confirmed dementia diagnoses? Arch Neurol. 2003;60(12):1696–1702. doi: 10.1001/archneur.60.12.1696. [DOI] [PubMed] [Google Scholar]
  • 210.Arriagada PV, Growdon JH, Hedley-Whyte ET, Hyman BT. Neurofibrillary tangles but not senile plaques parallel duration and severity of Alzheimer's disease. Neurology. 1992;42(3 Pt 1):631–639. doi: 10.1212/wnl.42.3.631. [DOI] [PubMed] [Google Scholar]
  • 211.Tariot PN, Aisen PS. Can lithium or valproate untie tangles in Alzheimer's disease? J Clin Psychiatry. 2009;70(6):919–921. doi: 10.4088/jcp.09com05331. [DOI] [PubMed] [Google Scholar]
  • 212.Roberson ED, Scearce-Levie K, Palop JJ, et al. Reducing endogenous tau ameliorates amyloid beta-induced deficits in an Alzheimer's disease mouse model. Science. 2007;316(5825):750–754. doi: 10.1126/science.1141736. [DOI] [PubMed] [Google Scholar]
  • 213.Santacruz K, Lewis J, Spires T, et al. Tau suppression in a neurodegenerative mouse model improves memory function. Science. 2005;309(5733):476–481. doi: 10.1126/science.1113694. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 214.Makrides V, Shen TE, Bhatia R, et al. Microtubule-dependent oligomerization of tau. Implications for physiological tau function and tauopathies. J Biol Chem. 2003;278(35):33298–33304. doi: 10.1074/jbc.M305207200. [DOI] [PubMed] [Google Scholar]
  • 215.Crowe A, Huang W, Ballatore C, et al. Identification of aminothienopyridazine inhibitors of tau assembly by quantitative high-throughput screening. Biochemistry. 2009;48(32):7732–7745. doi: 10.1021/bi9006435. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216.Wischik C, Staff R. Challenges in the Conduct of Disease-Modifying Trials in AD: Practical Experience from a Phase 2 Trial of Tau-Aggregation Inhibitor Therapy. J Nutr Health Aging. 2009;13(4):367–369. doi: 10.1007/s12603-009-0046-5. [DOI] [PubMed] [Google Scholar]
  • 217.Wischik CM, Edwards PC, Lai RY, Roth M, Harrington CR. Selective inhibition of Alzheimer disease-like tau aggregation by phenothiazines. Proc Natl Acad Sci U S A. 1996;93(20):11213–11218. doi: 10.1073/pnas.93.20.11213. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 218.Steinhilb ML, Dias-Santagata D, Fulga TA, Felch DL, Feany MB. Tau phosphorylation sites work in concert to promote neurotoxicity in vivo. Mol Biol Cell. 2007;18(12):5060–5068. doi: 10.1091/mbc.E07-04-0327. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 219.Schneider A, Mandelkow E. Tau-based treatment strategies in neurodegenerative diseases. Neurotherapeutics. 2008;5(3):443–457. doi: 10.1016/j.nurt.2008.05.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 220.Illenberger S, Zheng-Fischhofer Q, Preuss U, et al. The endogenous and cell cycle-dependent phosphorylation of tau protein in living cells: implications for Alzheimer's disease. Mol Biol Cell. 1998;9(6):1495–1512. doi: 10.1091/mbc.9.6.1495. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 221.Gong CX, Liu F, Grundke-Iqbal I, Iqbal K. Post-translational modifications of tau protein in Alzheimer's disease. J Neural Transm. 2005;112(6):813–838. doi: 10.1007/s00702-004-0221-0. [DOI] [PubMed] [Google Scholar]
  • 222.Wen Y, Planel E, Herman M, et al. Interplay between cyclin-dependent kinase 5 and glycogen synthase kinase 3 beta mediated by neuregulin signaling leads to differential effects on tau phosphorylation and amyloid precursor protein processing. J Neurosci. 2008;28(10):2624–2632. doi: 10.1523/JNEUROSCI.5245-07.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 223.Plattner F, Angelo M, Giese KP. The roles of cyclin-dependent kinase 5 and glycogen synthase kinase 3 in tau hyperphosphorylation. J Biol Chem. 2006;281(35):25457–25465. doi: 10.1074/jbc.M603469200. [DOI] [PubMed] [Google Scholar]
  • 224.Leyhe T, Eschweiler GW, Stransky E, et al. Increase of BDNF serum concentration in lithium treated patients with early Alzheimer's disease. J Alzheimers Dis. 2009;16(3):649–656. doi: 10.3233/JAD-2009-1004. [DOI] [PubMed] [Google Scholar]
  • 225.Porsteinsson AP, Tariot PN, Jakimovich LJ, et al. Valproate therapy for agitation in dementia: open-label extension of a double-blind trial. Am J Geriatr Psychiatry. 2003;11(4):434–440. [PubMed] [Google Scholar]
  • 226.Porsteinsson AP, Tariot PN, Erb R, Gaile S. An open trial of valproate for agitation in geriatric neuropsychiatric disorders. Am J Geriatr Psychiatry. 1997;5(4):344–351. doi: 10.1097/00019442-199700540-00010. [DOI] [PubMed] [Google Scholar]
  • 227.Hampel H, Ewers M, Burger K, et al. Lithium trial in Alzheimer's disease: a randomized, single-blind, placebo-controlled, multicenter 10-week study. J Clin Psychiatry. 2009;70(6):922–931. [PubMed] [Google Scholar]
  • 228.Tariot PN AP, Cummings J, Jakimovich L, Schneider L, Thomas R, Becerra L, Loy R. The ADCS Valproate Neuroprotection Trial: Primary Efficacy and Safety Results. ICAD. 2009 O1-04-04. [Google Scholar]
  • 229.Arendash GW, Mori T, Cao C, et al. Caffeine Reverses Cognitive Impairment and Decreases Brain Amyloid-beta Levels in Aged Alzheimer's Disease Mice. J Alzheimers Dis. 2009;17(3):661–680. doi: 10.3233/JAD-2009-1087. [DOI] [PubMed] [Google Scholar]
  • 230.Maia L, de Mendonca A. Does caffeine intake protect from Alzheimer's disease? Eur J Neurol. 2002;9(4):377–382. doi: 10.1046/j.1468-1331.2002.00421.x. [DOI] [PubMed] [Google Scholar]
  • 231.Eskelinen MH, Ngandu T, Tuomilehto J, Soininen H, Kivipelto M. Midlife coffee and tea drinking and the risk of late-life dementia: a population-based CAIDE study. J Alzheimers Dis. 2009;16(1):85–91. doi: 10.3233/JAD-2009-0920. [DOI] [PubMed] [Google Scholar]
  • 232.Arendash GW, Schleif W, Rezai-Zadeh K, et al. Caffeine protects Alzheimer's mice against cognitive impairment and reduces brain beta-amyloid production. Neuroscience. 2006;142(4):941–952. doi: 10.1016/j.neuroscience.2006.07.021. [DOI] [PubMed] [Google Scholar]
  • 233.Vogelsberg-Ragaglia V, Schuck T, Trojanowski JQ, Lee VM. PP2A mRNA expression is quantitatively decreased in Alzheimer's disease hippocampus. Exp Neurol. 2001;168(2):402–412. doi: 10.1006/exnr.2001.7630. [DOI] [PubMed] [Google Scholar]
  • 234.Sontag E, Luangpirom A, Hladik C, et al. Altered expression levels of the protein phosphatase 2A ABalphaC enzyme are associated with Alzheimer disease pathology. J Neuropathol Exp Neurol. 2004;63(4):287–301. doi: 10.1093/jnen/63.4.287. [DOI] [PubMed] [Google Scholar]
  • 235.Gong CX, Shaikh S, Wang JZ, Zaidi T, Grundke-Iqbal I, Iqbal K. Phosphatase activity toward abnormally phosphorylated tau: decrease in Alzheimer disease brain. J Neurochem. 1995;65(2):732–738. doi: 10.1046/j.1471-4159.1995.65020732.x. [DOI] [PubMed] [Google Scholar]
  • 236.Kins S, Crameri A, Evans DR, Hemmings BA, Nitsch RM, Gotz J. Reduced protein phosphatase 2A activity induces hyperphosphorylation and altered compartmentalization of tau in transgenic mice. J Biol Chem. 2001;276(41):38193–38200. doi: 10.1074/jbc.M102621200. [DOI] [PubMed] [Google Scholar]
  • 237.Arendt T, Holzer M, Fruth R, Bruckner MK, Gartner U. Paired helical filament-like phosphorylation of tau, deposition of beta/A4-amyloid and memory impairment in rat induced by chronic inhibition of phosphatase 1 and 2A. Neuroscience. 1995;69(3):691–698. doi: 10.1016/0306-4522(95)00347-l. [DOI] [PubMed] [Google Scholar]
  • 238.Iqbal K, Alonso Adel C, El-Akkad E, et al. Significance and mechanism of Alzheimer neurofibrillary degeneration and therapeutic targets to inhibit this lesion. J Mol Neurosci. 2002;19(1-2):95–99. doi: 10.1007/s12031-002-0017-3. [DOI] [PubMed] [Google Scholar]
  • 239.Tanimukai H, Kudo T, Tanaka T, Grundke-Iqbal I, Iqbal K, Takeda M. Novel therapeutic strategies for neurodegenerative disease. Psychogeriatrics. 2009;9(2):103–109. doi: 10.1111/j.1479-8301.2009.00289.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 240.Gozes I, Morimoto BH, Tiong J, et al. NAP: research and development of a peptide derived from activity-dependent neuroprotective protein (ADNP). CNS Drug Rev. 2005;11(4):353–368. doi: 10.1111/j.1527-3458.2005.tb00053.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 241.Divinski I, Mittelman L, Gozes I. A femtomolar acting octapeptide interacts with tubulin and protects astrocytes against zinc intoxication. J Biol Chem. 2004;279(27):28531–28538. doi: 10.1074/jbc.M403197200. [DOI] [PubMed] [Google Scholar]
  • 242.Hashimoto Y, Kaneko Y, Tsukamoto E, et al. Molecular characterization of neurohybrid cell death induced by Alzheimer's amyloid-beta peptides via p75NTR/PLAIDD. J Neurochem. 2004;90(3):549–558. doi: 10.1111/j.1471-4159.2004.02513.x. [DOI] [PubMed] [Google Scholar]
  • 243.Gozes I, Divinski I. The femtomolar-acting NAP interacts with microtubules: Novel aspects of astrocyte protection. J Alzheimers Dis. 2004;6(6 Suppl):S37–41. doi: 10.3233/jad-2004-6s605. [DOI] [PubMed] [Google Scholar]
  • 244.Matsuoka Y, Gray AJ, Hirata-Fukae C, et al. 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(2):165–170. doi: 10.1385/jmn/31:02:165. [DOI] [PubMed] [Google Scholar]
  • 245.Matsuoka Y, Jouroukhin Y, Gray AJ, et al. A neuronal microtubule-interacting agent, NAPVSIPQ, reduces tau pathology and enhances cognitive function in a mouse model of Alzheimer's disease. J Pharmacol Exp Ther. 2008;325(1):146–153. doi: 10.1124/jpet.107.130526. [DOI] [PubMed] [Google Scholar]
  • 246.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(3):1091–1098. [PubMed] [Google Scholar]
  • 247**.Oddo S, Vasilevko V, Caccamo A, Kitazawa M, Cribbs DH, LaFerla FM. Reduction of soluble Abeta and tau, but not soluble Abeta alone, ameliorates cognitive decline in transgenic mice with plaques and tangles. J Biol Chem. 2006;281(51):39413–39423. doi: 10.1074/jbc.M608485200. [Transgenic animal model study that suggested reduction in soluble Aβ was necessary but not sufficient to improve cognitive function. Cognitive improvement could be achieved only with active vaccination that resulted in reduced soluble Aβ and reduced tau levels. Acute active and passive immunization reduced soluble Aβ, did not reduce tau, and did not result in cognitive improvement.] [DOI] [PubMed] [Google Scholar]
  • 248.Mendez MF. Frontotemporal dementia: therapeutic interventions. Front Neurol Neurosci. 2009;24:168–178. doi: 10.1159/000197896. [DOI] [PubMed] [Google Scholar]
  • 249.Strong MJ, Yang W, Strong WL, Leystra-Lantz C, Jaffe H, Pant HC. Tau protein hyperphosphorylation in sporadic ALS with cognitive impairment. Neurology. 2006;66(11):1770–1771. doi: 10.1212/01.wnl.0000218161.15834.db. [DOI] [PubMed] [Google Scholar]
  • 250.McGeer PL, Schulzer M, McGeer EG. Arthritis and anti-inflammatory agents as possible protective factors for Alzheimer's disease: a review of 17 epidemiologic studies. Neurology. 1996;47(2):425–432. doi: 10.1212/wnl.47.2.425. [DOI] [PubMed] [Google Scholar]
  • 251.Szekely CA, Breitner JC, Fitzpatrick AL, et al. NSAID use and dementia risk in the Cardiovascular Health Study: role of APOE and NSAID type. Neurology. 2008;70(1):17–24. doi: 10.1212/01.wnl.0000284596.95156.48. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 252.Dufouil C, Richard F, Fievet N, et al. APOE genotype, cholesterol level, lipid-lowering treatment, and dementia: the Three-City Study. Neurology. 2005;64(9):1531–1538. doi: 10.1212/01.WNL.0000160114.42643.31. [DOI] [PubMed] [Google Scholar]
  • 253.Haag MD, Hofman A, Koudstaal PJ, Stricker BH, Breteler MM. Statins are associated with a reduced risk of Alzheimer disease regardless of lipophilicity. The Rotterdam Study. J Neurol Neurosurg Psychiatry. 2009;80(1):13–17. doi: 10.1136/jnnp.2008.150433. [DOI] [PubMed] [Google Scholar]
  • 254.Sparks DL, Scheff SW, Hunsaker JC, 3rd, Liu H, Landers T, Gross DR. Induction of Alzheimer-like beta-amyloid immunoreactivity in the brains of rabbits with dietary cholesterol. Exp Neurol. 1994;126(1):88–94. doi: 10.1006/exnr.1994.1044. [DOI] [PubMed] [Google Scholar]
  • 255.Aisen PS, Schafer KA, Grundman M, et al. Effects of rofecoxib or naproxen vs placebo on Alzheimer disease progression: a randomized controlled trial. Jama. 2003;289(21):2819–2826. doi: 10.1001/jama.289.21.2819. [DOI] [PubMed] [Google Scholar]
  • 256.Martin BK, Szekely C, Brandt J, et al. Cognitive function over time in the Alzheimer's Disease Anti-inflammatory Prevention Trial (ADAPT): results of a randomized, controlled trial of naproxen and celecoxib. Arch Neurol. 2008;65(7):896–905. doi: 10.1001/archneur.2008.65.7.nct70006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 257.Group AR, Lyketsos CG, Breitner JC, et al. Naproxen and celecoxib do not prevent AD in early results from a randomized controlled trial. Neurology. 2007;68(21):1800–1808. doi: 10.1212/01.wnl.0000260269.93245.d2. [DOI] [PubMed] [Google Scholar]
  • 258.Meinert CL, Breitner JC. Chronic disease long-term drug prevention trials: lessons from the Alzheimer's Disease Anti-inflammatory Prevention Trial (ADAPT). Alzheimers Dement. 2008;4(1 Suppl 1):S7–S14. doi: 10.1016/j.jalz.2007.10.007. [DOI] [PubMed] [Google Scholar]
  • 259.Arvanitakis Z, Schneider JA, Wilson RS, et al. Statins, incident Alzheimer disease, change in cognitive function, and neuropathology. Neurology. 2008;70(19 Pt 2):1795–1802. doi: 10.1212/01.wnl.0000288181.00826.63. [DOI] [PubMed] [Google Scholar]
  • 260.Rea TD, Breitner JC, Psaty BM, et al. Statin use and the risk of incident dementia: the Cardiovascular Health Study. Arch Neurol. 2005;62(7):1047–1051. doi: 10.1001/archneur.62.7.1047. [DOI] [PubMed] [Google Scholar]
  • 261.Li G, Larson EB, Sonnen JA, et al. Statin therapy is associated with reduced neuropathologic changes of Alzheimer disease. Neurology. 2007;69(9):878–885. doi: 10.1212/01.wnl.0000277657.95487.1c. [DOI] [PubMed] [Google Scholar]
  • 262.Sparks DL, Sabbagh MN, Connor DJ, et al. Atorvastatin for the treatment of mild to moderate Alzheimer disease: preliminary results. Arch Neurol. 2005;62(5):753–757. doi: 10.1001/archneur.62.5.753. [DOI] [PubMed] [Google Scholar]
  • 263.Jones RW, Kivipelto M, Feldman H, et al. The Atorvastatin/Donepezil in Alzheimer's Disease Study (LEADe): design and baseline characteristics. Alzheimers Dement. 2008;4(2):145–153. doi: 10.1016/j.jalz.2008.02.001. [DOI] [PubMed] [Google Scholar]
  • 264.Feldman H, et al. The LEADe Study: A Randomized, Controlled Trial Investigating the Effect of Atorvastatin on Cognitive and Global Function in Patients With Mild-to-Moderate Alzheimer's Disease Receiving Background Therapy of Donepezil. AAN. 2008 Abstract LBS.005. [Google Scholar]
  • 265.Green DR, Reed JC. Mitochondria and apoptosis. Science. 1998;281(5381):1309–1312. doi: 10.1126/science.281.5381.1309. [DOI] [PubMed] [Google Scholar]
  • 266.Spierings D, McStay G, Saleh M, et al. Connected to death: the (unexpurgated) mitochondrial pathway of apoptosis. Science. 2005;310(5745):66–67. doi: 10.1126/science.1117105. [DOI] [PubMed] [Google Scholar]
  • 267.Bachurin SO, Shevtsova EP, Kireeva EG, Oxenkrug GF, Sablin SO. Mitochondria as a target for neurotoxins and neuroprotective agents. Ann N Y Acad Sci. 2003;993:334–344. doi: 10.1111/j.1749-6632.2003.tb07541.x. discussion 345-339. [DOI] [PubMed] [Google Scholar]
  • 268.Beal MF. Mitochondria take center stage in aging and neurodegeneration. Ann Neurol. 2005;58(4):495–505. doi: 10.1002/ana.20624. [DOI] [PubMed] [Google Scholar]
  • 269.Caspersen C, Wang N, Yao J, et al. Mitochondrial Abeta: a potential focal point for neuronal metabolic dysfunction in Alzheimer's disease. FASEB J. 2005;19(14):2040–2041. doi: 10.1096/fj.05-3735fje. [DOI] [PubMed] [Google Scholar]
  • 270.Sirk D, Zhu Z, Wadia JS, et al. Chronic exposure to sub-lethal beta-amyloid (Abeta) inhibits the import of nuclear-encoded proteins to mitochondria in differentiated PC12 cells. J Neurochem. 2007;103(5):1989–2003. doi: 10.1111/j.1471-4159.2007.04907.x. [DOI] [PubMed] [Google Scholar]
  • 271.Doody RS, Gavrilova SI, Sano M, et al. Effect of dimebon on cognition, activities of daily living, behaviour, and global function in patients with mild-to-moderate Alzheimer's disease: a randomised, double-blind, placebo-controlled study. Lancet. 2008;372(9634):207–215. doi: 10.1016/S0140-6736(08)61074-0. [DOI] [PubMed] [Google Scholar]
  • 272.Cummings JL, et al. 18-month data from an open-label extension of a one-year controlled trial of dimebon in mild-to-moderate Alzheimer's disease. ICAD. 2008:P4–334. [Google Scholar]
  • 273.Urdinguio RG, Sanchez-Mut JV, Esteller M. Epigenetic mechanisms in neurological diseases: genes, syndromes, and therapies. Lancet Neurol. 2009;8(11):1056–1072. doi: 10.1016/S1474-4422(09)70262-5. [DOI] [PubMed] [Google Scholar]
  • 274.Antonello M, Rotili D, Valente S, Kazantsev AG. Histone Deacetylase Inhibitors and Neurodegenerative Disorders: Holding the Promise. Curr Pharm Des. 2009 doi: 10.2174/138161209789649349. [DOI] [PubMed] [Google Scholar]
  • 275.Fischer A, Sananbenesi F, Wang X, Dobbin M, Tsai LH. Recovery of learning and memory is associated with chromatin remodelling. Nature. 2007;447(7141):178–182. doi: 10.1038/nature05772. [DOI] [PubMed] [Google Scholar]
  • 276.Guan JS, Haggarty SJ, Giacometti E, et al. HDAC2 negatively regulates memory formation and synaptic plasticity. Nature. 2009;459(7243):55–60. doi: 10.1038/nature07925. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 277.Green KN, Steffan JS, Martinez-Coria H, et al. Nicotinamide restores cognition in Alzheimer's disease transgenic mice via a mechanism involving sirtuin inhibition and selective reduction of Thr231-phosphotau. J Neurosci. 2008;28(45):11500–11510. doi: 10.1523/JNEUROSCI.3203-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 278.Fumagalli F, Molteni R, Calabrese F, Maj PF, Racagni G, Riva MA. Neurotrophic factors in neurodegenerative disorders : potential for therapy. CNS Drugs. 2008;22(12):1005–1019. doi: 10.2165/0023210-200822120-00004. [DOI] [PubMed] [Google Scholar]
  • 279.Apfel SC. Is the therapeutic application of neurotrophic factors dead? Ann Neurol. 2002;51(1):8–11. doi: 10.1002/ana.10099. [DOI] [PubMed] [Google Scholar]
  • 280.Blesch A. Neurotrophic factors in neurodegeneration. Brain Pathol. 2006;16(4):295–303. doi: 10.1111/j.1750-3639.2006.00036.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 281.Eriksdotter Jonhagen M, Nordberg A, Amberla K, et al. Intracerebroventricular infusion of nerve growth factor in three patients with Alzheimer's disease. Dement Geriatr Cogn Disord. 1998;9(5):246–257. doi: 10.1159/000017069. [DOI] [PubMed] [Google Scholar]
  • 282.Tuszynski MH. Growth-factor gene therapy for neurodegenerative disorders. Lancet Neurol. 2002;1(1):51–57. doi: 10.1016/s1474-4422(02)00006-6. [DOI] [PubMed] [Google Scholar]
  • 283.Whitehouse PJ, Price DL, Struble RG, Clark AW, Coyle JT, Delon MR. Alzheimer's disease and senile dementia: loss of neurons in the basal forebrain. Science. 1982;215(4537):1237–1239. doi: 10.1126/science.7058341. [DOI] [PubMed] [Google Scholar]
  • 284.Tuszynski MH, Thal L, Pay M, et al. A phase 1 clinical trial of nerve growth factor gene therapy for Alzheimer disease. Nat Med. 2005;11(5):551–555. doi: 10.1038/nm1239. [DOI] [PubMed] [Google Scholar]
  • 285.Emerich DF, Winn SR, Harper J, Hammang JP, Baetge EE, Kordower JH. Implants of polymer-encapsulated human NGF-secreting cells in the nonhuman primate: rescue and sprouting of degenerating cholinergic basal forebrain neurons. J Comp Neurol. 1994;349(1):148–164. doi: 10.1002/cne.903490110. [DOI] [PubMed] [Google Scholar]
  • 286.Chen KS, Gage FH. Somatic gene transfer of NGF to the aged brain: behavioral and morphological amelioration. J Neurosci. 1995;15(4):2819–2825. doi: 10.1523/JNEUROSCI.15-04-02819.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 287.Nagahara AH, Bernot T, Moseanko R, et al. Long-term reversal of cholinergic neuronal decline in aged non-human primates by lentiviral NGF gene delivery. Exp Neurol. 2009;215(1):153–159. doi: 10.1016/j.expneurol.2008.10.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 288.Nagahara AH, Merrill DA, Coppola G, et al. Neuroprotective effects of brain-derived neurotrophic factor in rodent and primate models of Alzheimer's disease. Nat Med. 2009;15(3):331–337. doi: 10.1038/nm.1912. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 289.Bishop KM, Hofer EK, Mehta A, et al. Therapeutic potential of CERE-110 (AAV2-NGF): targeted, stable, and sustained NGF delivery and trophic activity on rodent basal forebrain cholinergic neurons. Exp Neurol. 2008;211(2):574–584. doi: 10.1016/j.expneurol.2008.03.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 290.Alvarez XA, Cacabelos R, Laredo M, et al. A 24-week, double-blind, placebo-controlled study of three dosages of Cerebrolysin in patients with mild to moderate Alzheimer's disease. Eur J Neurol. 2006;13(1):43–54. doi: 10.1111/j.1468-1331.2006.01222.x. [DOI] [PubMed] [Google Scholar]
  • 291.Brookmeyer R, Gray S, Kawas C. Projections of Alzheimer's disease in the United States and the public health impact of delaying disease onset. Am J Public Health. 1998;88(9):1337–1342. doi: 10.2105/ajph.88.9.1337. [DOI] [PMC free article] [PubMed] [Google Scholar]

Web References

  1. www.clinicaltrials.gov.

RESOURCES