Next Article in Journal
Role of Mitochondrial Dysfunction in Degenerative Brain Diseases, an Overview
Previous Article in Journal
Blacks’ Diminished Return of Education Attainment on Subjective Health; Mediating Effect of Income
Previous Article in Special Issue
Clearing Extracellular Alpha-Synuclein from Cerebrospinal Fluid: A New Therapeutic Strategy in Parkinson’s Disease
 
 
Editorial published on 11 September 2020, see Brain Sci. 2020, 10(9), 633.
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Neurodegenerative Diseases: Regenerative Mechanisms and Novel Therapeutic Approaches

1
Center for Translational Neuromedicine, University of Rochester, NY 14642, USA
2
Department of Animal Sciences, Quaid-i-Azam University, Islamabad 45320, Pakistan
*
Authors to whom correspondence should be addressed.
Brain Sci. 2018, 8(9), 177; https://doi.org/10.3390/brainsci8090177
Submission received: 13 July 2018 / Revised: 3 September 2018 / Accepted: 12 September 2018 / Published: 15 September 2018
(This article belongs to the Special Issue Pathogenesis and Treatment of Neurodegenerative Diseases)

Abstract

:
Regeneration refers to regrowth of tissue in the central nervous system. It includes generation of new neurons, glia, myelin, and synapses, as well as the regaining of essential functions: sensory, motor, emotional and cognitive abilities. Unfortunately, regeneration within the nervous system is very slow compared to other body systems. This relative slowness is attributed to increased vulnerability to irreversible cellular insults and the loss of function due to the very long lifespan of neurons, the stretch of cells and cytoplasm over several dozens of inches throughout the body, insufficiency of the tissue-level waste removal system, and minimal neural cell proliferation/self-renewal capacity. In this context, the current review summarized the most common features of major neurodegenerative disorders; their causes and consequences and proposed novel therapeutic approaches.

1. Introduction

Regeneration processes within the nervous system are referred to as neuroregeneration. It includes, but is not limited to, the generation of new neurons, axons, glia, and synapses. It was not considered possible until a couple of decades ago, when the discovery of neural precursor cells in the sub-ventricular zone (SVZ) and other regions shattered the dogma [1,2,3,4]. Neuroregeneration can also be defined as the progressive structural and functional recovery of the damaged nervous system over time. Damage to the central nervous system (CNS) is attributed to cell death, axonal regeneration failure, demyelination, and overall neuronal structural and functional deficits. All these conditions—partially or wholly, solitary or combined, genetic or acquired, known or unknown in origin—are manifested in specific neurological disorders, collectively termed as neurodegenerative disorders. These disorders jeopardize the normal functioning of the brain and lead to the progressive decline or even the complete loss of sensory, motor, and cognitive function. Examples include, but are not limited to Alzheimer’s disease (AD), Huntington’s disease (HD), Parkinson’s disease (PD), and multiple sclerosis (MS).
Importantly, neurodegenerative diseases manifest in an abnormal buildup of proteins in the brain/tissue, i.e., β-amyloid in AD, misfolded Huntington protein in HD, aggregation of ubiquitinated proteins in amyotrophic lateral sclerosis [5], Tau and β-amyloid accumulation in MS plaques [6], α-synuclein accumulation in PD, and Tau neurofibrillary tangles in traumatic brain injuries [7]. Evidence suggests that the spread of misfolded protein from cell-to-cell significantly contributes to the progression of disease [8]. Moreover, in many cases, these misfolded proteins invade healthy brain tissue when two of these affected cells are placed together [8].
Given the widespread pervasiveness of potential neurodegeneration within the brain, many structures and regions are impaired. Consequently, synaptic insufficiency, massive cell death, inflammation, temporary or permanent loss of various bodily actions like coordinated motion (ataxias) and different cognitive skills like memory (dementia), decision-making skills, talking, breathing, and heart function, also become prominent features of neuropathology [9,10,11,12,13]. Whilst there are currently no cures for neurodegenerative diseases, particularly in advanced stages, developing therapeutic techniques is of the utmost importance to overcome physiological and cognitive deficits [14]. Recently, scientific literature has devoted countless studies to providing insight on novel therapy techniques to counteract and prevent the damaging effects of neurodegenerative diseases [15,16,17,18], particularly through ameliorating immune system and hormonal therapy, i.e., testosterone, estrogens, GH/IGF, etc. [19,20,21]. As research progresses, a greater understanding of the mechanisms that contribute to the progressive degeneration of neurons and their connections and the ultimate loss of cognitive and motor skills, could lead to more effective therapeutic techniques in the near future.

2. Causes and Consequences of Neurodegeneration

All neurodegenerative diseases affect different regions of the brain, whilst exhibiting distinctive and apparent characteristics at the phenotypic level, i.e., progressive loss of sensory-motor and cognitive functions [22,23], but overall they share similar etiology at the cellular and molecular level [24,25,26]. Critical analysis of the similarities between these disorders offers the potential for therapeutic advancements, which could tackle many of these diseases simultaneously if we clearly understand the commonalities existing between various neurodegenerative disorders [27,28]. In this respect, neurodegeneration can be seen at different levels of neuronal circuitry, ranging from disturbance of intra-cellular protein molecules to inter-cellular disturbance of tissue and overall systems.
Out of many different types of neurodegenerative diseases, Alzheimer’s (AD), Parkinson’s (PD), Huntington’s diseases (HD), and multiple sclerosis (MS), are the most commonly occurring forms. AD is the leading cause of dementia worldwide, causing the inability of an individual to perform everyday activities. An estimated 5.4 million Americans have AD, including approximately 200,000 aged <65 years, which comprises the younger-onset AD population. Statistics also show that every passing 68 seconds adds another patient of AD [29]. It starts as mild memory loss with symptoms worsening over time, and the affected person forgets how to perform basic daily activities like combing their hair and brushing their teeth. Over time, they become unable to recognize family members and need permanent care, which becomes a burden on society. The buildup of β-amyloid protein and intracellular aggregation of tau protein are the noxious etiological agents, which might trigger synaptopathies, glial inflammation, and eventual neuronal death in the cerebral cortex, sub-cortical regions, temporal and parietal lobes, and cingulate gyrus, observed in AD [30,31,32,33,34,35].
PD is characterized as a movement disorder, with an incidence of 0.3% in industrialized countries (Parkinson’s Association). It is caused by a decrease in brain dopamine (DA) levels. Death of DA producing neurons in substantia nigra is triggered by the intracellular accumulation of protein α-synuclein bound to ubiquitin complex [36]. These protein aggregates form cytoplasmic inclusions, commonly known as Lewy’s bodies, which play a significant role in familial and sporadic cases of Parkinson’s disease [37,38,39,40,41,42,43]. Symptoms begin as shaking of hands, arms, legs, and neck muscles. With time, severe ataxia develops, and the person fails to perform his everyday tasks [44,45].
Similarly, HD is also caused by the intracellular accumulation of aggregates of a mutant Huntington’s protein, resulting in brain cell apoptosis mainly in the striatum [46,47,48,49,50,51]. HD is a genetic disorder exhibiting movement problems and ataxia like PD after middle age, and progressive neuronal decline leads to the dementia-like AD [52,53,54].
AD, PD, and HD are primarily classified as proteinopathies, meaning they are associated with aggregation of misfolded proteins. AD and PD, both late-onset, are progressive diseases linked to the intracellular accumulation of toxic protein aggregates. Although, these diseases exhibit distinctive and apparent characteristics at the phenotypic level, at the subcellular level, they have a common etiology. In AD, neuronal death is seen in the amygdala, cortex, and hippocampus [9]. While in PD, substantia nigra shows neuronal loss leading to dopamine deficiency in the striatum [55].
In contrast to the above mentioned neuronal disorders, Multiple Sclerosis (MS) is a glial disorder. It involves massive damage to myelinated fibers through autoimmune reaction, causing axonal injury and further loss of neuronal communication mostly in the white matter tracts, the basal ganglia, and the brain stem [56,57,58,59,60,61,62,63]. This disease leads to a variety of physical, mental, and psychiatric problems [64,65,66,67]. In comparison to AD and PD, HD and MS are early in onset. AD, PD, and HD all have a strong genetic predisposition [68,69,70,71,72,73,74,75,76,77], where potentially the mutation is hereditarily passed down family lineage. In the AD, genetic heritability ranges from 49–79%, based on twin and family studies statistics [78]. In PD, 5–10% of those afflicted with the disease showed a mutation in different genes [79], leaving the individuals at greater risk to develop the symptoms. Similarly, as a purely a genetic disorder, HD is caused by trinucleotide repeat expansions in the Huntington protein [46]. Though unlike these genetically predisposed diseases, MS is not considered a hereditary disease. However, some genetic variations are believed to increase the risk of developing the disorder [80,81,82,83,84].
As discussed earlier, AD and PD share a phenotypic feature: Dementia. Dementia is an overarching term used to describe a group of symptoms, which result in a severe long-term decline in cognitive function significant enough to affect daily function. It can result from a number of complex disorders that damage the brain, including but not limited to AD, vascular dementia, frontotemporal dementia, dementia with Lewy bodies, and PD. Typical symptoms of dementia can include deficits in memory and language, impaired visuospatial skills, loss of executive function and attention, as well as behavioral disturbances [85]. Interestingly, autopsy analysis of a significant number of individuals diagnosed with AD or PD does not show the hallmark pathological feature of protein aggregate deposition [86,87]. Vascular dementia (caused by stroke) is also reported to be frequently occurring with AD, and this phenomenon could aggravate the symptoms of dementia [87]. Depending on the phenotypic features, vascular dementia can be misdiagnosed as AD or PD [88]. Overlapping pathological features of these diseases add further to diagnostic complexity. Data from studies investigating the link between neuropathology and molecular genetics have shown that phenotypic symptoms are not always highly associated with the underlying etiological changes, and they can be triggered by a number of other factors like experience, cognitive reserve, and epigenetics [89].
Focusing the neuropathological markers of dementia and defining it as a biological construct can lead to the discovery of novel genetic variants, the development of new peripheral biomarkers, as well as the identification of individuals at higher risk of the disease before the appearance of clinical symptoms [90]. This will also become an increasingly important issue as new drug treatments are developed [91].

3. Intra-Cellular Signaling Mechanisms

Compelling evidence indicates that the mTOR signaling pathway is involved in disease progress, aging, and regeneration. mTOR is a serine/threonine kinase, which in conjugation with other proteins makes two complexes: mTOR complex 1 (mTORC1) and mTOR complex 2 (mTORC2). Both complexes are phosphorylated by AKT dependent Pi3K, in which localization of both is solely cytoplasmic, but the field of operation is completely different. mTORC1 having raptor as its integral component promotes protein synthesis, ribosome biogenesis, proliferation, migration, and differentiation, by stimulating S6K1 and inhibiting 4EBP1 and elF4E. Whilst mTORC2 having rictor as its integral component promotes cell survival, cell cycle progression, and actin remodeling by its actions through PKC and SGK1. Studies suggest that mTOR promotes many of the processes which are impaired in HD, AD, PD, and MS [92,93,94]. However, the strategy of stimulating mTOR for a gain of function varies from disease-to-disease and case-by-case.
For example, mTOR signaling is perturbed in PD [95], which is mainly attributed to a lack of L-DOPA resulting in the degeneration of the basal ganglia’s dopamine. However, treatment of L-DOPA nonspecifically activates mTOR; this over-activation produces symptoms of dyskinesia, which can be prevented by simultaneous treatment with rapamycin [96]. In mouse models of Parkinson’s disease, treatment of L-DOPA activates mTOR complex one, which is involved in neural synaptic rehabilitation directed toward the basal ganglia, proving to be a promising therapeutic approach for future clinical trials of PD [97].
Similarly, AD postmortem tissue analysis suggests a hyperactivation of mTOR/AKT/Pi3K, and increased levels of its downstream targets p70S6K and 4EBP1. This mTOR hyper-activation is significantly correlated with increases in the levels of beta-amyloid plaques [94]. Moreover, pre-clinical studies suggest that mTOR activation enhances Aβ generation and deposition by modulating amyloid precursor protein (APP) metabolism and upregulating β- and γ-secretases [98], while its inhibition by rapamycin ameliorates AD like conditions [93,99,100]. In HD, the mTOR pathway is hijacked by abnormally accumulated Huntington protein, which increases the mTORC1 activity [101], and results in massive damage possibly through inhibition of autophagy [102].
On the other hand, multiple sclerosis is a complex disorder where there is massive damage governed by inflammation [103,104], and in this disorder, mTOR activity is high and needs to be toned down [105,106,107]. Pre-clinical studies suggest that mTOR inactivation prevents the development of experimental autoimmune encephalomyelitis (EAE) in mice [108]. On the other hand, regeneration involving proliferation of oligodendrocyte, and subsequent differentiation clearly demands increased mTOR activity [109,110,111].
Beyond the mTOR complex implicated in neurodegenerative diseases, other signaling pathways have been found to be disrupted in disease pathology, which includes but is not limited to bone morphogenetic protein (BMP), mitogen activated protein kinase (MAPK), wnt/β catenin signaling, etc. [112,113]. Interruption of the BMP signaling pathway has been associated with failed synapse formation and maintenance in several neurodegenerative diseases. Decreased signaling has been noted in HD, while abnormally increased signaling has been found in MS [114]. In the case of Huntington patients, high levels of a protein thought down-regulate BMP signaling had been found in high concentrations [115]. It is thought that this excess is caused by the defected axonal transport. Axonal transport defects have been correlated with these BMP signaling deficiencies, such as in AD, even decades before official diagnosis [116]. Similarly, in AD patients, defected axonal transport has been observed along with the synaptic dysfunction in the inferior temporal gyrus, which has been associated with the loss of cognitive function [117]. Since the BMP signaling pathway maintains synapse function, the disruption of this pathway leads to an onslaught of negative consequences. BMP pathway is implicated in MS with increased signaling not only in mouse models of MS, but in MS patients [118]. Inhibition of the wnt/β-catenin pathway, which is essential for blood brain barrier formation, exacerbates MS like conditions in the animal model experimental autoimmune encephalomyelitis [113]. Overactivated MAPK signaling in AD increases neuronal apoptosis, as well as buildup of β-amyloid, through increased expression of β and γ secretases [113,119]; whilst it promotes neurogenesis in the mouse model of PD [120].

4. Current Treatment Paradigm

Current treatment strategies of neurodegenerative diseases merely target a small subset of the population and focus on symptomatic relief only, without altering disease progression (Summarized in Figure 1). This results in permanent disability or death of those afflicted. Presently, the Food and Drug Administration (FDA) has approved acetylcholine esterase inhibitors [Donepezil (Aricept), Rivastigmine (Exelon)], to be used as palliative treatment (see Figure 1 for action in different brain regions).
These drugs reduce the symptoms and slow down the progression of the disease [121,122], but they are not useful in long-term treatment of the disease. PD is treated with Levodopa in combination with carbidopa (Sinemet), which crosses the blood-brain barrier and gets converted to dopamine after decarboxylation. It restores dopamine levels in the substantia nigra, and ameliorates all of the clinical features of Parkinsonism for the first few years, but loses efficacy on prolonged use [123]. Dopamine agonists [Pergolide (Permax), Bromocriptine (Parlodel)] are also in practice, but they cause a variety of adverse effects, including cardiovascular and endocrinological problems [124,125]. In contrast to PD, HD is caused by overactivity in dopaminergic nigrostriatal pathways. As a result, its treatment uses drugs that impair the dopaminergic transmission, either by depleting central monoamines [e.g., Reserpine (Serpadil)] or by blocking dopamine receptors [e.g., phenothiazines (Haldol, Trilafon)] [126]. MS is treated with many immuno-suppressers, which help to speed up recovery from relapse and slow down the progression of the disease. Therapies include MS relapse prevention by prednisone to reduce inflammation, Ocrelizumab for primary progressive MS, and a few other drugs for relapse re-emitting MS, including: beta-interferon (immunomodulatory), Ocrelizumab (neutralizing antibodies), glatiramer acetate, alemtuzumab and mitoxantrone (immuno-suppressor), Tysabri, and natalizumab, which provoke immune cells to enter into the brain [127,128,129,130] (Figure 1).

5. Future Strategies

Despite advancements in the fields of molecular biology, genetics, and pharmaceutical sciences, current biomedical science is a long way from the ultimate goal of identifying the risk factors, specific diagnostic tools, and appropriate strategies for the effective treatment of these neurodegenerative disorders. With an increasing burden on society, a need for newer treatment strategies is also in increasing demand. Current treatment approach mainly relies upon the survival of affected neurons and slowing the disease progression. This writing is based on some novel approaches (summarized in Table 1), which can promote the longevity of afflicted neurons and might slow down the degeneration process, improving the recovery rates.

5.1. Inhibiting and Disaggregating Protein Aggregates

A typical mammalian cell contains around 30,000 proteins, which are made up of linear chains of amino acids [131], and are responsible for all necessary biological actions. After transcription and translation, these proteins undergo post-translational modifications (mainly folding) to achieve the functional native form. Proper protein folding is under the control of the Protein Quality Control System (PQC) present inside the cell, which has evolved over an extended period [132]. Despite the activity of this system, many proteins fail to achieve normal functional folding state, because of multiple reasons like aging [133], oxidative stress [134], gene mutations [135], altered cellular temperature [136,137], and pH [138], etc.
Molecular chaperones act on these misfolded proteins to partially unfold them, which forms specific intermediates that might rearrange by themselves to build oligomeric aggregates that finally convert to amyloid fibrils [139,140]. In case of failure of chaperons, the lysosomal degradation system comes to the cellular rescue and targets these misfolded proteins. However, if this surveillance system also fails to cope with the situation, then the accumulation of these protein aggregates leads to the appearance of common neurodegenerative symptoms, caused by the amyloidosis of the CNS [141,142].
Short peptides, called synthetic mini-chaperons consisting of a recognition motif of a misfolded protein have shown some therapeutic potential in amyloid diseases. An example is heat shock protein (Hsp), a major molecular chaperone. Mammalian Hsps have been classified into families by their molecular weight, i.e., Hsp110, Hsp90, Hsp70, Hsp60, Hsp40, and Hsp27 [143,144]. Modification of the heat shock protein Hsp104, which upon interaction with other Hsps like Hsp110, Hsfp70, Hsp40, etc., very efficiently disaggregates non-desired proteins in yeast cells [145]. These peptides are designed so that they bind with the protein fibril at one end, but the other end does not contain any binding site [131], thus blocking the aggregation process. Moreover, modifications (N-methylation) of the amino acid backbone blocks intermolecular H-bonding, offering a steric hindrance in the formation of amyloid [146,147]. Promising effects of Hsp104 have been shown to dissolve various amyloid conformations and deposits of pre-amyloid oligomers. However, its application in humans requires a very high level dose in the present conformity, which seems utterly impractical. Similarly, several in-vivo and in-vitro studies have reported a neuroprotective role of Hsp70 and related compounds [148], by enhancing clearance of Aβ aggregates [149], and by restoring tau homeostasis [150]. This action of Hsp70 is attributable to upregulation of the insulin degradation enzyme (IDE) and TGF-β1expression [148]. Based on these molecular mechanisms, four major therapeutic strategies are proposed.

5.1.1. Induction of Endogenous Hsp70 Production

Different compounds like galandamycin [151], geranylgeranylacetone [152], and celastrol [153], have been reported to induce endogenous Hsp70 production. These compounds give new leads for therapeutic strategy development related to neurodegenerative diseases.

5.1.2. Application of Exogenous Hsp70

In a recent study, intranasal administration of Hsp70 was reported to improve cognitive deficits in AD like pathology in the mice model [154], possibly by reducing the production of reactive oxygen species [155]. However, this effect is not long lasting. Developing more effective dosage forms can be a potential therapeutic strategy for enhanced effect of exogenous Hsp70.

5.1.3. Constitutive Expression of Hsp70

Two cytosolic variants of Hsp70: Hsc70 and Hsp72, were previously reported to influence tau confirmation, degradation, and aggregation [156]. These variants are reported to have opposing effects on tau clearance. Hsp72 is reported to improve tau clearance, whilst Hsc70 is found to delay the clearance of tau protein aggregates [157]. Developing strategies to enhance Hsp72 expression selectively, and decrease Hsc70 expression, can prove to be an essential strategy to remove toxic protein aggregates and improve the pathological symptoms.

5.1.4. Inhibition of Hsp70 ATPase

Recently, selective Hsp70 ATPase inhibitor YM-08, has been found to reduce the level of pathogenic tau levels in brain cells, and retained its affinity in vitro [158]. Finding more chemicals like this can prove to be a crucial and highly selective therapeutic technique for AD.
An alternative strategy is to increase endogenous production of Hsps. Studies show that cellular stress is the primary factor that upregulates Hsp, through activation of heat shock factor 1 (HSF-1). In healthy cells, Hsp most importantly, Hsp90 binds with HSF-1 and keeps it in an inactive form. While under stress, these repressors are occupied by misfolded proteins, thus leaving the HSF-1 un-repressed. Which leads to HSF-1 translocation to the nucleus, trimerization, phosphorylation at Ser4-419 residue and chromatin binding [159,160]. The exact details of this pathway and involvement of other protein complexes are not entirely known. However, a few studies show that Ral1 binding protein 1, tubulin, and P23 are very influential, and could be a potential drug target to increase endogenous production [161,162]. Collectively, these studies suggest a neuroprotective role of these molecular chaperones, but further efforts are needed to design a more efficient sequence to cope with dose associated problems, and the successful application in neurodegenerative disorders in humans to target β-Amyloid and Tau [145,163,164]. Late-onset neurodegenerative diseases, including PD and HD, are linked with the formation of intracellular toxic aggregates of different proteins. Therefore, understanding the factors that regulate the homeostatic levels of these proteins at synthesis and degradation stages is crucial [165].
After the aggregates have been formed, the normal function and structure of a protein can be restored by dissolving the aggregate. Protein disaggregates possess an innate property of breaking down these aggregates, and have shown therapeutic potential in the mitigation of the toxicity caused by these aggregates [166]. They dismantle self-templating amyloids and prion structures, which form neurotoxic oligomers. Different Hsp has shown potential in dissolving the protein aggregates, in metazoan and eukaryotic cells [167]. One example is eukaryotic Hsp104, an ATPase, but its activity is suboptimum. It can be therapeutically engineered to achieve an optimum level of therapeutic activity by induction of some mutations [167]. One variant of Hsp104 has shown therapeutic efficacy in dissolving α-synuclein aggregates in PD models [168].
HD is characterized by the polyQ, a residue that is known to cause protein misfolding in mHTT mutant Huntington [169]. Studies have found that Hsp70 targets the buildup of mHTT by binding to the proteins and halting further misfolding [170], preventing further neurotoxicity. Chaperones belonging to Hsp70 have also been found to specifically degrade polyQ proteins characteristic in HD, which has been observed both in vivo mice studies and in flies [171]. Stimulation of Hsp70 and the chaperones within that family by plant extracts, i.e., polyphenol curcumin or epigallocatechin gallate etc., can alleviate the cognitive and motor deficits that result from the neuronal death/damage caused by misfolded mHTT [172,173].
PD, on the other hand, is characterized by a buildup of α-synuclein misfolded protein aggregates. Hsp70, another molecular chaperone, identifies the α-synuclein by its hydrophobic region and binds to the protein, halting further misfolding and stimulating refolding [174]. Studies have shown that the absence of Hsp70 in disease models accelerates neurodegeneration in PD models, demonstrating the active role that this molecular chaperone plays in inhibiting further neuronal damage [175].
Certain Hsp, while still active in identifying and refolding protein aggregates, have also been found to assist in the buildup of misfolded proteins by protecting the neurotoxic proteins. Hsp90 has been found to support neurotoxic proteins often [176]. A pharmaceutical drug, geldanamycin has been found to inhibit protein buildup in PD dopaminergic murine models [177]. However, because geldanamycin cannot pass through the blood-brain barrier, other drugs must be developed to target HSF1, a transcription factor that stimulates the transcription of other HSPs that can effectively dissolve misfolded protein aggregates, as demonstrated in in-vitro studies.
The specialized roles of Hsp and their affiliated chaperones, have been found to specifically target certain types of protein aggregates in different neurodegenerative diseases. The identification of the specific Hsp and their correlated chaperones that provide the most significant assistance in targeting particular protein aggregates (mHTT, α-synuclein, etc.) for each neurodegenerative disorder and the development of drugs stimulating these heat shock proteins can prove beneficial for future therapeutic techniques [178]. Additionally, identifying Hsp that support neurotoxic buildup allows for targeted inhibition of these specific proteins, whilst stimulating the transcription of other useful Hsp and their relevant chaperones.
Combining protein disaggregation with protein degradation can be helpful in the elimination of toxic loss/gain of functional proteins, and reclamation of neuronal health.

5.2. Immuno-Modulation

Philosophically, neuroinflammation can be regarded as beneficial for the neuronal tissue to clean debris; but the scenario is more complicated when inflammatory mediators stay in the tissue longer than they are needed. It starts a vicious cycle of inflammation, which ultimately results in neuronal death and neurodegeneration. Elevated plasma and Cerebrospinal Fluid (CSF) concentrations of proinflammatory cytokines, such as IL-6, TNFα, IL-1β, IL-2, IL-6, and cyclooxygenase-1/2 have been found in PD, AD, and HD [179,180,181].
Microglia are present in the HD brain even before the onset of symptoms [182]. Detection of activated microglia and astrocytes have been found in HD patients up to 15 years before the symptomatic onset, and PET scans have indicated a correlation between microglia accumulation and HD pathology [183,184]. The activated microglia undergo morphological changes, such as an increased appearance of cytokines, specifically cytokine IL-6, in which increased levels are associated with increased severity of the functional decline in HD patients. Activated microglia, also undergo a morphological change in an ameboid fashion, meaning that the microglia take on an immune response with an increase of phagocytic activity, whilst also responding to damaged tissue. Therefore, detecting activated microglia even before the onset of HD symptoms is imperative in predicting the disorder’s severity [182,185,186]. Activated microglia can be classified as M1 and M2, based on the release of cytokines. Classically, M1 microglia release pro-inflammatory cytokines (IL-1β, IL-6, IL-8, etc.) and chemokines (CCL2, CCL20 and CXCL-1), whilst M2 microglia release anti-inflammatory cytokines and growth factors (IL-10, TGFβ, CD206 etc.); the latter initiates tissue repair and regeneration [187,188]. Post mortem studies suggest that HD patients have a significantly lower number of M2 microglia [188,189].
The central immune activation reflects a low-grade immune response in the periphery [190]. Furthermore, monocytes isolated from HD gene carriers, expressing the mutant Huntingtin protein, show hyperactivity to lipopolysaccharide stimulation [191]. Taken together, a hyperactive immune system proves to be an important feature of HD pathogenesis, and developing a strategy that deals with this hyperreactivity, using immunomodulators, is a potential disease-modifying avenue in HD treatment. Moreover, CD8+ and CD4+ peripheral lymphocytes have also been found in the substantia nigra of post mortem brains of PD patients [192]. Observational evaluation of large cohorts of patients, suggests that the use of non-steroidal anti-inflammatory drugs reduces the risk of PD [193].
Similarly, minocycline a potent anti-inflammatory drug which blocks activation of NAPDH oxidase and microglial activation, has also proven very effective in rodent models of PD [194]. Natural anti-inflammatory compounds like resveratrol [195,196], tanshinone [197,198], and silymarin [199], have shown therapeutic promise in animal models of PD, by downregulating the glial cell activation and pro-inflammatory cytokine release. Recently, sargramostim has shown therapeutic promise in a randomized, double-blind clinical trial [200]. In a recent in vitro study, laquinimod (an oral immunomodulator) demonstrated a dampening effect on the proinflammatory cytokine release from activated monocytes, and brings them down to base line activity [201].
Most of these inflammatory effects are mediated by microglia, where shutting down the microglia response altogether may even be more detrimental. It has been proposed that therapies should shift microglia response from pro-inflammatory M1 to the anti-inflammatory M2 Phase. The tone of microglia, i.e., pro-inflammatory to anti-inflammatory, can be managed by selectively improving the infiltration of regulatory T cells (Treg), producing IL-10 and BDNF [202,203].
Use of monoclonal antibodies against the α -synuclein, reduced the levels of protein propagation and demonstrated improved PD-like pathologies, ameliorated dopaminergic neuronal cell loss, and attenuated motor deficits in mouse models of the disease [204,205], whilst reducing amyloid formation [204,205]. It should also be noted that autoantibodies that can detect and degrade amyloidogenic proteins, like Ab, α-synuclein, and PrP are already present in the human serum [166,206].
Selective monoclonal antibodies to pathological proteins offer a novel avenue for the treatment of neurodegenerative diseases. Several monoclonal antibodies have been reported, which specifically bind to prefibrillar oligomers and not to amyloid fibrils, monomer, or natively folded proteins. Like the polyclonal antisera, the individual monoclonals recognize generic epitopes that do not depend on a specific linear amino acid sequence, but they display distinct preferences for different subsets of prefibrillar oligomers [207,208]. It is reported that autoantibodies that can detect and degrade amyloidogenic proteins, like Ab, α-synuclein, and PrP are already present in the human serum [166,206]. Table 2 highlights the antibodies currently undergoing clinical trials, for the treatment of proteinopathies.

5.3. Stimulating Autophagy

Autophagy, a cellular phenomenon responsible for removal of protein aggregates and dysfunctional mitochondria, is critical for the survival of long living, non-dividing cells, such as neurons and myocytes [209,210], as these cells are expected to accumulate unfolded or misfolded proteins in a significant amount because of their longer lifespan.
Initially, autophagy was thought to be inactive in neurons, as autophagosomes were found to be either completely absent from the neuronal cells or were found very rarely. Later, it was found that neuronal cells have a very effective lysosomal system, which is responsible for removal of autophagosomes rapidly, and inhibition of lysosomal activity leads to accumulation of autophagosomes [211]. It is not surprising that the mutation of gene-regulating autophagy causes neurodegenerative disorders, including AD, ALS, and PD [212]. Neurons have a very large expanse of axonal and dendritic cytoplasm. A long life span and limited regeneration make it very easy for neurons to gather large debris of dysfunctional organelles and waste proteins, which cannot be diluted because of limited cell division. Typically, phagocytosis starts with a change in phosphorylation of Unc51 by mTOR complex 11, followed by waste substance enwrapping by membrane-bound vacuole and fusion with lysosomes [212]. Various cellular stress signals, including decreased concentrations of amino acids, growth factors, hypoxia, and mechanical damages, suppress neuronal functions and plasticity through suppression of autophagy [213,214,215]. Flies and mice with autophagy gene (e.g., ATG) mutation, had a visible decline in neuronal health [216,217]. Specifically, the accumulation of aggregates of ubiquitin and faulty mitochondria was noticed in these cells [216,217], which are the hallmarks of neurodegeneration. These findings highlight the importance of autophagy in the maintenance of neuronal health.
Drug-induced autophagy is a promising regulatory approach to mediating the removal of abnormal proteins. Recently non-toxic, small molecules that induce autophagy in neurons have shown research potential as a therapeutic technique in animal models. Methyl-4-phenylpyridinium (MPP+) exposure has been found to induce apoptosis, in both in vivo and in vitro dopaminergic neurons in mouse Parkinson’s Models [218]. MPP+ is a neurotoxin that accumulates in dopaminergic neurons and organelles, with the toxin disrupting the complex I of the electron transport chain of mitochondria [219]. This inhibition induced by MPP+ leads to neuronal apoptosis.
Similarly, latrepirdine, with the drug-patented name of Dimedbon, is an antihistamine that demonstrates promising results in mouse AD trials [220]. Latrepirdine regulates and protects the amyloid-β precursor protein (APP) associated with the buildup of Aβ proteins in Alzheimer’s. Meta-analysis showed a modest effect on behavior, but it did not show any beneficial effect on cognition in mild to moderate AD patients [221].
Whilst Lithium has been classically a major treatment of affective disorders, recent experimental and clinical studies have shown neuroprotective effects beneficial for the treatment of neurodegenerative pathologies, such as the regulation of autophagy and the synthesis of neurotrophic factors [222]. Drugs such as rapamycin and calcium channel blockers have also shown potential in stimulating the autophagic process in various studies done in Drosophila [223,224] and mouse models [102,225], where these showed an improved hepatic autophagic function in obese mice and increased clearance of mutant huntingtin protein. Rapamycin, an FDA approved drug, has been used therapeutically to reduce Aβ -induced cognitive deficits of Alzheimer’s [226]. It is also reported to reduce the risk of the AD in Type 2 Diabetes Mellitus (T2DM) rats by inhibiting the activation of AMPK-mTOR signaling pathway [227]. Mammalian rapamycin (mTOR) regulates protein synthesis and degradation in the mTOR complexes 1 and 2 [228]. Signaling of mTOR has been found to induce apoptotic pathways by forming autophagic vacuoles, ameliorating tau and Aβ in mouse models.
Metformin has also shown beneficial effects in animal models of different neurodegenerative diseases like HD [229]. It induces the autophagic process by an AMPK-dependent mechanism. It also dephosphorylates neurofibrillary tangles of tau in AD clinical trials, by inducing protein phosphatase 2A (PP2A) activity in transgenic mice models [220]. As the major tau phosphatase in the brain, the induction of PP2A, which the drug Metformin targets, is vital to its therapeutic potential for neurodegenerative pathology [230]. Similarly, nilotinib has shown therapeutic promise in the improvement of PD profile by stimulating the autophagic process, inhibiting the tyrosine kinase activity in PD [231] and AD [232] mouse models. Continuing studies which highlight the further exploration of autophagy-improving chemicals can help in mitigating pervasive neurodegenerative symptoms (Summary in Table 3).

5.4. Clearance by Glymphatic System

Cellular debris and waste products are cleared from the body’s tissue by the lymphatic system. Unfortunately, the brain lacks a fully functional lymphatic system [233]. However, Nedergaard and co-workers discovered a parallel system in the brain, which drains neuronal metabolites by convective flow of cerebral spinal fluid through perivascular spaces [234,235,236]. The clearance pathway largely depends upon astrocyte-specific water channels AQP4; efficiency of this mechanism decreases with age, with an increase in AQP4 depolarization [237,238]. Deletion of these water channels which facilitate the convective fluid transport of CSF into the brain parenchyma and interstitial fluid back towards perivenous spaces [239], impair the normal influx/efflux of fluids. Moreover, this glia dependent lymphatic system analog—referred to as the Glymphatic System—is active only during sleep [234]. This identifies the importance of sleep, as well as the convective flow of CNS fluid, for proper and efficient removal of waste materials from the brain.
More recent findings suggest that impairment of the glymphatic system is a potential risk factor for the development of neurodegenerative disorders, which are particularly associated with the buildup of abnormal protein as in AD [240,241,242]. Accumulation of β-amyloid protein in the intestinal space and along the vasculature, is a characteristic feature of AD. Importantly, its production and aggregation increase with age [243]. As β-amyloid accumulates, the glymphatic flow becomes disrupted, and waste clearance deteriorates.
Beyond just the waste clearance itself, the other functions of the glymphatic system have yet to be fully understood. Since the glymphatic system is thought to distribute lipids via the blood-brain barrier, it is possible that this system mediates growth factors and other neuromodulators which contribute to functional recovery in neurodegenerative diseases. Perhaps the glymphatic system distributes the molecules involved in neurogenesis, axonal growth, and neurotrophic factors. Thus, therapies to combat the negative domino effects of age on the glymphatic system, used in conjunction with the current understanding of the glymphatic system, should undergo further investigation.

5.5. Neurogenesis and Neurotrophic Factors

Despite neurodegenerative damage to the brain, a degree of functional recovery is made possible through specific cellular and structural mechanisms that occur soon after the onset of an injury or cell death. Discoveries over the last couple of decades have refined the conception of the presence of neural stem cells within the brain. Neural stem cells are distributed throughout the brain, their presence and proliferation have convincingly been shown in the sub-ventricular zone, dentate gyrus of the hippocampus, and rostral migratory stream [244,245,246,247]. However, activation of these cells to participate in repair demands a very organized orchestration of events, which can only be visible during embryonic, early postnatal periods, and to a lesser extent during/after injury or brain insult. It involves active cell-to-cell communication, the presence of neurotrophic factors, and the activation of intra-cellular signaling mechanisms.
Neurotrophic factors are imperative for recovery in neurodegenerative diseases. Often, neurodegenerative diseases result in dysregulation of neurotrophic factors [248,249,250,251], molecules that are specific to types of neurons and aid in neuron function and survival, specifically in proliferation, differentiation, and growth. Without functional regulation of these neurotrophic factors, neurons and glia tend to change shape and decrease in number, in affected regions of the brain [252,253,254,255,256]. Therefore, the mechanisms that regulate these neurotrophic factors must be galvanized with the onset of neurodegenerative diseases.
Neurotrophic factors play two specific roles in the development and maintenance of the nervous system. During development, neurotrophic factors stimulate synaptic connections between neurons, and support axonal growth [257,258]. Then throughout adulthood, the neurotrophic factors maintain these synaptic connections and inhibit nerve cell apoptosis, regulating brain function [259,260]. Briefly, neurotrophic families consist of nerve growth factors (NGF), brain-derived neurotrophic factors (BDNF), ciliary neurotrophic factors (CNF), glial cell line-derived neurotrophic factors, Ephrins, endothelial growth factors (EGF) and transforming growth factors (TGF), neurotrophin 3 (NT3), and neurotrophin 4 (NT4). NGF and BDNF are particularly important in potential neurodegenerative disease therapies because, though they bind to different tyrosine receptor kinases, both utilize a similar pathway to promote cell survival through inhibition of apoptotic signals, and promote tissue growth by stimulating proliferation [261,262].
Since NGF and BDNF play important roles in the development and maintenance of neuron function and survival, the link between these neurotrophic factors and neurodegenerative diseases is unsurprising. Alterations of NGF and BDNF in the brain, and the disrupted binding of NGF and BDNF to their kinase receptors, are both linked to neurodegenerative diseases [263]. For example, studies have shown that a decreased level of NGF in Alzheimer’s patients leads to cellular death [264,265]. Similarly, studies have shown a decrease of BDNF in vulnerable brain areas, such as the substantia nigra in Parkinson’s patients [266,267]. In these areas of decreased neurotrophic expression, synaptic connections degenerate, leading to a cascade of synaptic transmission malfunction throughout this brain area without the proper maintenance that NGF and BDNF normally perform. As a result, neuron size and number begin to decrease, followed by a loss in neuronal function.
Therefore, it seems only logical that the most straightforward therapeutic approach to combat this loss of neuronal function is to increase the neurotrophins in the given degenerating area of the brain; however, as therapeutic based research has exhibited, certain challenges arise when invasively delivering neurotrophins. Since neurotrophins are polar and relatively large molecules, they invariably cross the highly selective blood-brain barrier [268]. Furthermore, selectively controlling delivery of neurotrophins to only damaged neurons, rather than functional ones, is nearly impossible.
Currently, research has provided researchers with several approaches to target neurotrophic deficits in the brain, despite these addressed challenges. These approaches include direct intracerebroventricular injection, viral vector-mediated gene delivery injection, and neurotrophin mimetics [263]. Whilst all these approaches exhibit potential, each approach comes with its share of side-effects. To avoid the negative side-effects of direct invasive injection into damaged brain areas, small molecules that mimic neurotrophins in its binding to specific tyrosine kinases is another viable option. These molecules are not neurotrophins themselves, but activate the same receptors, resulting in the same response as the binding of a neurotrophin to the receptor [61,269].
Other than neurotrophic factors, neural development in health and disease largely depends upon the presence of neurosteroids. The term was coined by French physiologist Etienne Baulieu, who demonstrated for the first time that the brain was capable of local synthesis of steroid hormones [270], to identify them from those produces by gonads separately. Being a very small molecular entity and non-polar nature, they cross the blood-brain barrier without any difficulty. Any change or deficiency within the neural tissue is readily compensated by circulating levels in the bloodstream. Deficiency in levels of these neurosteroids has been clinically validated in AD [271,272,273,274,275], PD [276,277], HD [278], and MS [279,280,281]; hormonal replacement therapy depicts beneficial effects in AD [276,282,283,284,285,286], PD [276,277,287,288,289], HD, and MS [256,290,291,292,293].
Stimulation of androgen receptors using neuro-active testosterone has beneficial influence on experimental autoimmune encephalomyelitis, a widely used disease model for the immune-mediated and inflammatory aspects of MS [294]. These neuroprotective effects are attributed to its immunomodulary and anti-inflammatory actions [295,296,297]. Recent clinical trials in relapse remitting MS patients, suggested an improved cognitive performance [256,292]. A detailed study of testosterone and its synthetic analogue in acute and chronic demyelination models, suggests that these effects are dependent upon the presence of neural androgen receptors [19]. Mice lacking an androgen receptor in neural tissue failed to recover and remyelinate their white matter tracts, even though they were administered with testosterone or its synthetic analog [19]. Similarly, the beneficial effects of testosterone and other neuroactive steroids (i.e., progesterone, estrogens etc.) have been documented for other neurological disorders like AD [276,282,283,284,285,286], PD [276,277,287,288,289], HD, and MS [256,290,291,292,293]. These neuro-regenerative effects of these neuroactive steroids most likely go through classical nuclear receptors [19,298,299,300,301,302].
Similarly, several studies showed the neuroprotective effect of estradiol in MS, AD, PD, and HD [263,303,304,305,306,307], whilst the lack of estrogens within the brain makes it more prone to neurodegenerative disorders. Administration of exogenous estradiol enhances neurogenesis by a mechanism that requires both α and β subtypes of estrogen receptors [308,309]. Additionally, estradiol promotes the migration of newly generated neurons towards the damaged brain regions [310]. Dual actions of estradiol: neuroprotective and anti-inflammatory [311,312], make it an ideal candidate for down-regulating the expression of pro-inflammatory cytokines, thus facilitating neuroprotection [313,314].
Presence and amelioration of the above mentioned neurotrophic alone are not enough. Regeneration still depends upon intracellular signaling. Typically, cells sense extra-cellular environmental changes through several membrane receptors (i.e., ligand-gated ion channels and receptors, enzyme-coupled receptors, GPCRs, and integrin receptors, etc.). These receptors are highly responsive to a very small change in extracellular environmental cues. Once a message is received, it is transferred across the membrane, and it involves a series/cascade of interactions between cytoplasmic proteins, including protein kinase B (AKT) [315,316,317], PI3K [318,319], Integrin-linked kinase (ILK) [320,321,322], microtubule-associated protein/ERK (p44/42) [323,324], mammalian target of rapamycin (mTOR) [325,326,327], and many more, which either translocate to the nucleus or activate chromatin binding proteins resulting in cell growth, proliferation, or differentiation (Figure 2).

5.6. Insulin and Neurodegeneration

In the central nervous system, insulin signaling promotes differentiation, proliferation, and neurite growth, and possesses neuroprotective and anti-apoptotic activity [328]. Insulin also modulates the structure and function of synapses, neurons, and neuronal circuits while playing a role in learning, cognition, and memory [328]. Neurochemical changes in neurodegenerative diseases may relate to abnormalities in glucose metabolism in vivo [329].
Brain insulin resistance partly drives cognitive impairment and AD in humans with peripheral insulin resistance diseases, including diabetes mellitus and metabolic syndromes [330,331]. Brain insulin resistance is also associated with increased levels of phospho-Tau and Aβ42 [332]. Experimental models have shown that brain insulin deficiency impairs learning and memory [333]. In early or intermediate stages of the AD, brain and CSF levels of insulin are found to be decreased [334], whilst Aβ42 and advanced glycation end-products are increased [334,335]. Insulin administration is found to improve working memory and cognition [336,337], and enhances Aβ42 clearance from the brain [337].
Activation of the insulin receptor (IR) leads to recruitment of insulin response substrate (IRS1 or 2) intracellularly [338,339], mediating the downstream signaling cascade. Insulin signaling is regulated by normal inhibition of IRS1 via MAPks [339]. The activity of these kinases can be modulated by extracellular factors like inflammatory mediators [340], oxidative stress [341], and Aβ oligomers [342,343]. Later, IRS-1 pSer616 and IRS-1 pSer636/639 were identified as putative biomarkers for brain insulin resistance in AD, and were reported to correlate positively with Aβ oligomer levels and negatively with cognitive function [344].
Dense representation of insulin receptors is reported on the dopaminergic neurons of the substantia nigra pars compacta [345]. Similarly, loss of insulin receptor immunoreactivity and messenger RNA in the substantia nigra pars compacta of patients with PD correlates with loss of tyrosine hydroxylase messenger RNA (i.e., the rate-limiting enzyme in dopamine synthesis) [346,347]. Abnormal glucose utilization has been noted in the brains of PD patients [348,349]. Abnormal glucose metabolism and increased frequency of T2DM has also been reported in patients with HD [350,351]. However, this link has not been studied intensively, and the underlying molecular mechanisms are not reported.
Metformin, a common FDA approved the antidiabetic drug, is reported to inhibit glucose production by the liver and to increase the glucose uptake in peripheral tissues, thereby lowering the blood glucose levels [352,353]. Until now, only a few animal studies have assessed the effect of metformin on cognitive decline, and the results differ [354,355,356,357]. Whilst the results of clinical studies have shown a positive effect of metformin on cognition and neuropathological features [358,359,360,361], a greater understanding of metformin’s mechanism of action and development of more agents with similar activity may help to create novel therapeutic targets for these notorious diseases.

5.7. Cholinergic System in AD

Recent data suggest that in contrast to early-onset AD, late-onset AD is a complex polygenic disease, which implicates atypical interaction between different molecular pathways. Clinically, the disease expression highlights the malfunction and ultimate collapse of both structural and neurochemical neuronal networks, including the cholinergic system [362]. Cholinergic synapses are ubiquitous in the human central nervous system. The significance of cholinergic transmission in higher brain functions like learning and memory is highlighted by the presence of high-density cholinergic synapses in the thalamus, striatum, limbic system, and neocortex [362].
Acetylcholine, an important neurotransmitter of the central nervous system, shows activity through the whole cortex, basal ganglia, and basal forebrain [363]. Human autopsy studies have shown that cholinergic loss is based on the degeneration of nucleus basalis of Meynert (NBM) cholinergic neurons, and of the axons they project to the cerebral cortex [364,365]. This cholinergic hypothesis of AD pathophysiology revolutionized the field of AD research, by carrying it from the traditional therapeutics to the contemporary concept of remedies based on synaptic neurotransmission. Three major discoveries led to the development of this hypothesis: (1) Discovery of depleted presynaptic cholinergic markers in the cerebral cortex [364,365]; (2) Discovery of the NBM in the basal forebrain as the major source of cortical cholinergic innervation, which undergoes severe degeneration in AD [366,367]; and (3) The demonstration that cholinergic antagonists weaken the memory, whilst agonists have the opposite effect [368]. This hypothesis received fascinating recognition when significant symptomatic improvement was induced by cholinesterase inhibitor therapy in AD patients [369], while anticholinergic drugs were reported to negatively affect human learning and memory [368,370,371].
Current therapeutic strategy in the management of AD, is based on the reclamation of cholinergic function by hindering the breakdown of acetylcholine [372,373] using cholinesterase inhibitors. Contemporary FDA-approved cholinesterase inhibitors in clinical practice are donepezil, rivastigmine, and galantamine. These compounds are reported to significantly improve cognition and some other behavioral features related to AD [373]. A meta-analysis of 26 studies of donepezil, rivastigmine, and galantamine reported a modest, but clinically significant, overall advantage of these drugs for stabilizing cognition, behavior, and global clinical change [374]. Early involvement of cholinergic system at preclinical stages of the disease was highlighted in different studies [375,376], which suggests that cholinomimetics can have a distinct role in disease prevention, along with other therapeutic agents. All this clinical and pathological data makes it very likely that cholinesterase inhibitors and discovery of other cholinomimetic agents can prove to be a very significant therapeutic intervention in the management of patients with AD.

6. Conclusions

In conclusion, the incidence of neurodegenerative disorders has been on the rise, and despite breakthrough discoveries, there remains an urgency from the patient’s perspective to search for and develop potential neuroprotective and neurorestorative therapeutics. Expansion of our understanding of intra, as well as inter, cellular signaling mechanisms both in health and disease, will greatly benefit our efforts to cure neurodegenerative disorders. Additionally, recapitulating the neuronal developmental paradigm in pathology and finding the means to create those conditions, including improved methods of drug delivery, would greatly enhance the chances of our success. Future research and clinical paradigms related to these notorious diseases may rely more heavily upon the ‘systems biology’ approach to these diseases, stressing the interaction of multiple factors such as genetic predisposition, stressors, inflammatory mechanisms, vascular insufficiency, dysregulation of protein aggregate formation, and clearance of neurofibrillary degeneration, cholinergic deficit, and other neurochemical anomalies. Therefore, despite substantial advances in the development of symptomatic treatments for neurodegenerative diseases, scientific efforts should not waiver, and perseverance is called for to attain this global goal.

Author Contributions

R.H. and M.S. outlined, wrote, and edited the review; whilst H.Z. and S.P. helped in writing and preparation of figures.

Funding

The research received no external funding.

Acknowledgments

We thank Dan Xue, Center for Translational Neuromedicine, University of Rochester, for assistance in graphics.

Conflicts of Interest

Authors declare no conflict of interest

References

  1. Gage, F.H.; Temple, S. Neural stem cells: Generating and regenerating the brain. Neuron 2013, 80, 588–601. [Google Scholar] [CrossRef] [PubMed]
  2. Malberg, J.E.; Eisch, A.J.; Nestler, E.J.; Duman, R.S. Chronic antidepressant treatment increases neurogenesis in adult rat hippocampus. J. Neurosci. Off. J. Soc. Neurosci. 2000, 20, 9104–9110. [Google Scholar] [CrossRef]
  3. Kuhn, H.G.; Dickinson-Anson, H.; Gage, F.H. Neurogenesis in the dentate gyrus of the adult rat: Age-related decrease of neuronal progenitor proliferation. J. Neurosci. Off. J. Soc. Neurosci. 1996, 16, 2027–2033. [Google Scholar] [CrossRef]
  4. Altman, J. Are new neurons formed in the brains of adult mammals? Science (New York) 1962, 135, 1127–1128. [Google Scholar] [CrossRef]
  5. Blokhuis, A.M.; Groen, E.J.N.; Koppers, M.; van den Berg, L.H.; Pasterkamp, R.J. Protein aggregation in amyotrophic lateral sclerosis. Acta Neuropathol. 2013, 125, 777–794. [Google Scholar] [CrossRef] [PubMed]
  6. David, M.A.; Tayebi, M. Detection of protein aggregates in brain and cerebrospinal fluid derived from multiple sclerosis patients. Front. Neurol. 2014, 5, 251. [Google Scholar] [CrossRef] [PubMed]
  7. Lucke-Wold, B.P.; Turner, R.C.; Logsdon, A.F.; Bailes, J.E.; Huber, J.D.; Rosen, C.L. Linking traumatic brain injury to chronic traumatic encephalopathy: Identification of potential mechanisms leading to neurofibrillary tangle development. J. Neurotrauma 2014, 31, 1129–1138. [Google Scholar] [CrossRef] [PubMed]
  8. Hatters, D.M. Protein misfolding inside cells: The case of huntingtin and Huntington’s disease. IUBMB Life 2008, 60, 724–728. [Google Scholar] [CrossRef] [PubMed]
  9. Martin, J.B. Molecular basis of the neurodegenerative disorders. N. Engl. J. Med. 1999, 340, 1970–1980. [Google Scholar] [CrossRef] [PubMed]
  10. Soto, C.; Estrada, L.D. Protein misfolding and neurodegeneration. Arch. Neurol. 2008, 65, 184–189. [Google Scholar] [CrossRef] [PubMed]
  11. Tuite, M.F.; Melki, R. Protein Misfolding and Aggregation in Ageing and Disease: Molecular Processes and Therapeutic Perspectives; Taylor & Francis: Abingdon, UK, 2007. [Google Scholar]
  12. Lepeta, K.; Lourenco, M.V.; Schweitzer, B.C.; Martino Adami, P.V.; Banerjee, P.; Catuara-Solarz, S.; de La Fuente Revenga, M.; Guillem, A.M.; Haidar, M.; Ijomone, O.M.; et al. Synaptopathies: Synaptic dysfunction in neurological disorders—A review from students to students. J. Neurochem. 2016, 138, 785–805. [Google Scholar] [CrossRef] [PubMed]
  13. Bae, J.R.; Kim, S.H. Synapses in neurodegenerative diseases. BMB Rep. 2017, 50, 237–246. [Google Scholar] [CrossRef] [PubMed]
  14. Kiaei, M. New hopes and challenges for treatment of neurodegenerative disorders: Great opportunities for young neuroscientists. Basic Clin. Neurosci. 2013, 4, 3–4. [Google Scholar] [PubMed]
  15. Duraes, F.; Pinto, M.; Sousa, E. Old drugs as new treatments for neurodegenerative diseases. Pharmaceuticals (Basel, Switzerland) 2018, 11, 44. [Google Scholar] [CrossRef] [PubMed]
  16. Sehgal, S.A.; Hammad, M.A.; Tahir, R.A.; Akram, H.N.; Ahmad, F. Current therapeutic molecules and targets in neurodegenerative diseases based on in silico drug design. Curr. Neuropharmacol. 2018, 16, 649–663. [Google Scholar] [CrossRef] [PubMed]
  17. Coppede, F. The potential of epigenetic therapies in neurodegenerative diseases. Front. Genet. 2014, 5, 220. [Google Scholar] [CrossRef] [PubMed]
  18. Amor, S.; Peferoen, L.A.; Vogel, D.Y.; Breur, M.; van der Valk, P.; Baker, D.; van Noort, J.M. Inflammation in neurodegenerative diseases—An update. Immunology 2014, 142, 151–166. [Google Scholar] [CrossRef] [PubMed]
  19. Hussain, R.; Ghoumari, A.M.; Bielecki, B.; Steibel, J.; Boehm, N.; Liere, P.; Macklin, W.B.; Kumar, N.; Habert, R.; Mhaouty-Kodja, S.; et al. The neural androgen receptor: A therapeutic target for myelin repair in chronic demyelination. Brain J. Neurol. 2013, 136, 132–146. [Google Scholar] [CrossRef] [PubMed]
  20. Behl, C. Oestrogen as a neuroprotective hormone. Nat. Rev. Neurosci. 2002, 3, 433–442. [Google Scholar] [CrossRef] [PubMed]
  21. Bondy, C.; Werner, H.; Roberts, C.T., Jr.; LeRoith, D. Cellular pattern of type-I insulin-like growth factor receptor gene expression during maturation of the rat brain: Comparison with insulin-like growth factors I and II. Neuroscience 1992, 46, 909–923. [Google Scholar] [CrossRef]
  22. Woolley, J.D.; Khan, B.K.; Murthy, N.K.; Miller, B.L.; Rankin, K.P. The diagnostic challenge of psychiatric symptoms in neurodegenerative disease: Rates of and risk factors for prior psychiatric diagnosis in patients with early neurodegenerative disease. J. Clin. Psychiatry 2011, 72, 126–133. [Google Scholar] [CrossRef] [PubMed]
  23. Montero-Odasso, M.; Pieruccini-Faria, F.; Bartha, R.; Black, S.E.; Finger, E.; Freedman, M.; Greenberg, B.; Grimes, D.A.; Hegele, R.A.; Hudson, C.; et al. Motor phenotype in neurodegenerative disorders: Gait and balance platform study design protocol for the ontario neurodegenerative research initiative (ONDRI). J. Alzheimer’s Dis. 2017, 59, 707–721. [Google Scholar] [CrossRef] [PubMed]
  24. Vadakkan, K.I. Neurodegenerative disorders share common features of “loss of function” states of a proposed mechanism of nervous system functions. Biomed. Pharmacother. Biomed. Pharmacother. 2016, 83, 412–430. [Google Scholar] [CrossRef] [PubMed]
  25. Hervas, R.; Oroz, J.; Galera-Prat, A.; Goni, O.; Valbuena, A.; Vera, A.M.; Gomez-Sicilia, A.; Losada-Urzaiz, F.; Uversky, V.N.; Menendez, M.; et al. Common features at the start of the neurodegeneration cascade. PLoS Boil. 2012, 10, e1001335. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Gitler, A.D.; Dhillon, P.; Shorter, J. Neurodegenerative disease: Models, mechanisms, and a new hope. Dis. Model Mech. 2017, 10, 499–502. [Google Scholar] [CrossRef] [PubMed]
  27. Rubinsztein, D.C. The roles of intracellular protein-degradation pathways in neurodegeneration. Nature 2006, 443, 780–786. [Google Scholar] [CrossRef] [PubMed]
  28. Thompson, L.M. Neurodegeneration: A question of balance. Nature 2008, 452, 707–708. [Google Scholar] [CrossRef] [PubMed]
  29. Alzheimer’s Association. 2012 Alzheimer’s disease facts and figures. Alzheimer’s Dement. J. Alzheimer’s Assoc. 2012, 8, 131–168. [Google Scholar] [CrossRef] [PubMed]
  30. Bloom, G.S. Amyloid-β and tau: The trigger and bullet in Alzheimer disease pathogenesis. JAMA Neurol. 2014, 71, 505–508. [Google Scholar] [CrossRef] [PubMed]
  31. Wenk, G.L. Neuropathologic changes in Alzheimer’s disease. J. Clin. Psychiatry 2003, 64, 7–10. [Google Scholar] [PubMed]
  32. Graham, W.V.; Bonito-Oliva, A.; Sakmar, T.P. Update on Alzheimer’s disease therapy and prevention strategies. Annu. Rev Med 2017, 68, 413–430. [Google Scholar] [CrossRef] [PubMed]
  33. Sanabria-Castro, A.; Alvarado-Echeverria, I.; Monge-Bonilla, C. Molecular pathogenesis of Alzheimer’s disease: An update. Ann. Neurosci. 2017, 24, 46–54. [Google Scholar] [CrossRef] [PubMed]
  34. Schneider, L. Alzheimer’s disease and other dementias: Update on research. Lancet Neurol. 2017, 16, 4–5. [Google Scholar] [CrossRef]
  35. Swerdlow, R.H.; Khan, S.M. The Alzheimer’s disease mitochondrial cascade hypothesis: An update. Exp. Neurol. 2009, 218, 308–315. [Google Scholar] [CrossRef] [PubMed]
  36. Ottolini, D.; Cali, T.; Szabo, I.; Brini, M. Alpha-synuclein at the intracellular and the extracellular side: Functional and dysfunctional implications. Boil. Chem. 2017, 398, 77–100. [Google Scholar] [CrossRef] [PubMed]
  37. Davie, C.A. A review of Parkinson’s disease. Br. Med. Bull. 2008, 86, 109–127. [Google Scholar] [CrossRef] [PubMed]
  38. Alkhuja, S. Parkinson disease: Research update and clinical management. South Med. J. 2013, 106, 334. [Google Scholar] [CrossRef] [PubMed]
  39. Frucht, S.J. Parkinson disease: An update. Neurologist 2004, 10, 185–194. [Google Scholar] [CrossRef] [PubMed]
  40. Rizek, P.; Kumar, N.; Jog, M.S. An update on the diagnosis and treatment of Parkinson disease. CMAJ 2016, 188, 1157–1165. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Chung, K.K.; Zhang, Y.; Lim, K.L.; Tanaka, Y.; Huang, H.; Gao, J.; Ross, C.A.; Dawson, V.L.; Dawson, T.M. Parkin ubiquitinates the alpha-synuclein-interacting protein, synphilin-1: Implications for Lewy-body formation in Parkinson disease. Nat. Med. 2001, 7, 1144–1150. [Google Scholar] [CrossRef] [PubMed]
  42. Frigerio, R.; Fujishiro, H.; Maraganore, D.M.; Klos, K.J.; DelleDonne, A.; Heckman, M.G.; Crook, J.E.; Josephs, K.A.; Parisi, J.E.; Boeve, B.F.; et al. Comparison of risk factor profiles in incidental Lewy body disease and Parkinson disease. Arch. Neurol. 2009, 66, 1114–1119. [Google Scholar] [CrossRef] [PubMed]
  43. Wood, H. Parkinson disease: Antibodies reveal age of Lewy pathology in PD. Nat. Rev. Neurol. 2017, 13, 318–319. [Google Scholar] [CrossRef] [PubMed]
  44. Schneider, J.S.; Sendek, S.; Yang, C. Relationship between motor symptoms, cognition, and demographic characteristics in treated mild/moderate Parkinson’s disease. PLoS ONE 2015, 10, e0123231. [Google Scholar] [CrossRef] [PubMed]
  45. Heinzel, S.; Bernhard, F.P.; Roeben, B.; Nussbaum, S.; Heger, T.; Martus, P.; Hobert, M.A.; Maetzler, W.; Berg, D. Progression markers of motor deficits in Parkinson’s disease: A biannual 4-year prospective study. Mov. Disord. Off. J. Mov. Disord. Soc. 2017, 32, 1254–1256. [Google Scholar] [CrossRef] [PubMed]
  46. Walker, F.O. Huntington’s disease. Lancet 2007, 369, 218–228. [Google Scholar] [CrossRef]
  47. Kaplan, A.; Stockwell, B.R. Therapeutic approaches to preventing cell death in Huntington disease. Prog. Neurobiol. 2012, 99, 262–280. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Messer, A.; McLear, J. The therapeutic potential of intrabodies in neurologic disorders: Focus on Huntington and Parkinson diseases. BioDrugs 2006, 20, 327–333. [Google Scholar] [CrossRef] [PubMed]
  49. Sciacca, G.; Cicchetti, F. Mutant huntingtin protein expression and blood-spinal cord barrier dysfunction in Huntington disease. Ann. Neurol. 2017, 82, 981–994. [Google Scholar] [CrossRef] [PubMed]
  50. Zhao, T.; Hong, Y.; Li, X.J.; Li, S.H. Subcellular clearance and accumulation of Huntington disease protein: A mini-review. Front. Mol. Neurosci. 2016, 9, 27. [Google Scholar] [CrossRef] [PubMed]
  51. Zheng, Z.; Diamond, M.I. Huntington disease and the huntingtin protein. Prog. Mol. Biol. Transl. Sci. 2012, 107, 189–214. [Google Scholar] [PubMed]
  52. Frank, S. Treatment of Huntington’s disease. Neurotherapeutics 2014, 11, 153–160. [Google Scholar] [CrossRef] [PubMed]
  53. Koutsis, G.; Panas, M.; Paraskevas, G.P.; Bougea, A.M.; Kladi, A.; Karadima, G.; Kapaki, E. From mild ataxia to Huntington disease phenocopy: The multiple faces of spinocerebellar ataxia 17. Case Rep. Neurol. Med. 2014, 2014, 643289. [Google Scholar] [CrossRef] [PubMed]
  54. Roxburgh, R.H.; Smith, C.O.; Lim, J.G.; Bachman, D.F.; Byrd, E.; Bird, T.D. The unique co-occurrence of spinocerebellar ataxia type 10 (SCA10) and Huntington disease. J. Neurol. Sci. 2013, 324, 176–178. [Google Scholar] [CrossRef] [PubMed]
  55. Hyman, B.T.; Van Hoesen, G.W. Neuron numbers in Alzheimer’s disease: Cell-specific pathology. Neurobiol. Aging 1987, 8, 555–556. [Google Scholar] [CrossRef]
  56. Bendfeldt, K.; Kappos, L.; Radue, E.W.; Borgwardt, S.J. Progression of gray matter atrophy and its association with white matter lesions in relapsing-remitting multiple sclerosis. J. Neurol. Sci. 2009, 285, 268–269. [Google Scholar] [CrossRef] [PubMed]
  57. Datta, G.; Colasanti, A.; Rabiner, E.A.; Gunn, R.N.; Malik, O.; Ciccarelli, O.; Nicholas, R.; Van Vlierberghe, E.; Van Hecke, W.; Searle, G.; et al. Neuroinflammation and its relationship to changes in brain volume and white matter lesions in multiple sclerosis. Brain J. Neurol. 2017, 140, 2927–2938. [Google Scholar] [CrossRef] [PubMed]
  58. Hacohen, Y.; Ciccarelli, O.; Hemingway, C. Abnormal white matter development in children with multiple sclerosis and monophasic acquired demyelination. Brain J. Neurol. 2017, 140, 1172–1174. [Google Scholar] [CrossRef] [PubMed]
  59. Lassmann, H. Mechanisms of white matter damage in multiple sclerosis. Glia 2014, 62, 1816–1830. [Google Scholar] [CrossRef] [PubMed]
  60. Longoni, G.; Brown, R.A.; MomayyezSiahkal, P.; Elliott, C.; Narayanan, S.; Bar-Or, A.; Marrie, R.A.; Yeh, E.A.; Filippi, M.; Banwell, B.; et al. White matter changes in paediatric multiple sclerosis and monophasic demyelinating disorders. Brain J. Neurol. 2017, 140, 1300–1315. [Google Scholar] [CrossRef] [PubMed]
  61. Josephy-Hernandez, S.; Jmaeff, S.; Pirvulescu, I.; Aboulkassim, T.; Saragovi, H.U. Neurotrophin receptor agonists and antagonists as therapeutic agents: An evolving paradigm. Neurobiol. Dis. 2017, 97, 139–155. [Google Scholar] [CrossRef] [PubMed]
  62. Moccia, M.; Quarantelli, M.; Lanzillo, R.; Cocozza, S.; Carotenuto, A.; Carotenuto, B.; Alfano, B.; Prinster, A.; Triassi, M.; Nardone, A.; et al. Grey:White matter ratio at diagnosis and the risk of 10-year multiple sclerosis progression. Eur. J. Neurol. 2017, 24, 195–204. [Google Scholar] [CrossRef] [PubMed]
  63. Suzuki, K.; Kamoshita, S.; Eto, Y.; Tourtellotte, W.W.; Gonatas, J.O. Myelin in multiple sclerosis. Composition of myelin from normal-appearing white matter. Arch. Neurol. 1973, 28, 293–297. [Google Scholar] [CrossRef] [PubMed]
  64. Chalah, M.A.; Ayache, S.S. Psychiatric event in multiple sclerosis: Could it be the tip of the iceberg? Rev. Bras. Psiquiatr. 2017, 39, 365–368. [Google Scholar] [CrossRef] [PubMed]
  65. de Cerqueira, A.C.; Semionato de Andrade, P.; Godoy Barreiros, J.M.; Teixeira, A.L.; Nardi, A.E. Psychiatric disorders in patients with multiple sclerosis. Compr. Psychiatry 2015, 63, 10–14. [Google Scholar] [CrossRef] [PubMed]
  66. Honer, W.G.; Hurwitz, T.; Li, D.K.; Palmer, M.; Paty, D.W. Temporal lobe involvement in multiple sclerosis patients with psychiatric disorders. Arch. Neurol. 1987, 44, 187–190. [Google Scholar] [CrossRef] [PubMed]
  67. Zambon, A.A.; Cecchetti, G.; Caso, F.; Santangelo, R.; Baldoli, C.; Natali Sora, M.G.; Comi, G.; Magnani, G.; Martinelli, V. Primary progressive multiple sclerosis presenting with severe predominant cognitive impairment and psychiatric symptoms: A challenging case. Mult. Scler. 2017, 23, 1558–1561. [Google Scholar] [CrossRef] [PubMed]
  68. Desikan, R.S.; Fan, C.C.; Wang, Y.; Schork, A.J.; Cabral, H.J.; Cupples, L.A.; Thompson, W.K.; Besser, L.; Kukull, W.A.; Holland, D.; et al. Genetic assessment of age-associated Alzheimer disease risk: Development and validation of a polygenic hazard score. PLoS Med. 2017, 14, e1002258. [Google Scholar] [CrossRef] [PubMed]
  69. Gaiteri, C.; Mostafavi, S.; Honey, C.J.; De Jager, P.L.; Bennett, D.A. Genetic variants in Alzheimer disease-molecular and brain network approaches. Nat. Rev. Neurol. 2016, 12, 413–427. [Google Scholar] [CrossRef] [PubMed]
  70. Li, Y.; Hollingworth, P.; Moore, P.; Foy, C.; Archer, N.; Powell, J.; Nowotny, P.; Holmans, P.; O’Donovan, M.; Tacey, K.; et al. Genetic association of the app binding protein 2 gene (APBB2) with late onset Alzheimer disease. Hum. Mutat. 2005, 25, 270–277. [Google Scholar] [CrossRef] [PubMed]
  71. Rosenberg, R.N.; Richter, R.W.; Risser, R.C.; Taubman, K.; Prado-Farmer, I.; Ebalo, E.; Posey, J.; Kingfisher, D.; Dean, D.; Weiner, M.F.; et al. Genetic factors for the development of Alzheimer disease in the Cherokee Indian. Arch. Neurol. 1996, 53, 997–1000. [Google Scholar] [CrossRef] [PubMed]
  72. Alcalay, R.N.; Caccappolo, E.; Mejia-Santana, H.; Tang, M.X.; Rosado, L.; Ross, B.M.; Verbitsky, M.; Kisselev, S.; Louis, E.D.; Comella, C.; et al. Frequency of known mutations in early-onset Parkinson disease: Implication for genetic counseling: The consortium on risk for early onset Parkinson disease study. Arch. Neurol. 2010, 67, 1116–1122. [Google Scholar] [CrossRef] [PubMed]
  73. Scott, W.K.; Staijich, J.M.; Yamaoka, L.H.; Speer, M.C.; Vance, J.M.; Roses, A.D.; Pericak-Vance, M.A. Genetic complexity and Parkinson’s disease. Deane laboratory Parkinson disease research group. Science (New York) 1997, 277, 387–388. [Google Scholar] [CrossRef]
  74. Verstraeten, A.; Theuns, J.; Van Broeckhoven, C. Progress in unraveling the genetic etiology of Parkinson disease in a genomic era. Trends Genet. 2015, 31, 140–149. [Google Scholar] [CrossRef] [PubMed]
  75. Bates, G.P. History of genetic disease: The molecular genetics of Huntington disease—A history. Nat. Rev. Genet. 2005, 6, 766–773. [Google Scholar] [CrossRef] [PubMed]
  76. Pericak-Vance, M.A.; Conneally, P.M.; Merritt, A.D.; Roos, R.; Norton, J.A., Jr.; Vance, J.M. Genetic linkage studies in Huntington disease. Cytogenet. Cell Genet. 1978, 22, 640–645. [Google Scholar] [CrossRef] [PubMed]
  77. Pulst, S.M. Neurodegenerative disease. Genetic discrimination in Huntington disease. Nat. Rev. Neurol. 2009, 5, 525–526. [Google Scholar] [CrossRef] [PubMed]
  78. Wilson, R.S.; Barral, S.; Lee, J.H.; Leurgans, S.E.; Foroud, T.M.; Sweet, R.A.; Graff-Radford, N.; Bird, T.D.; Mayeux, R.; Bennett, D.A. Heritability of different forms of memory in the late onset Alzheimer’s disease family study. J. Alzheimer’s Dis. 2011, 23, 249–255. [Google Scholar] [CrossRef] [PubMed]
  79. Lesage, S.; Brice, A. Parkinson’s disease: From monogenic forms to genetic susceptibility factors. Hum. Mol. Genet. 2009, 18, R48–R59. [Google Scholar] [CrossRef] [PubMed]
  80. Dyment, D.A.; Ebers, G.C.; Sadovnick, A.D. Genetics of multiple sclerosis. Lancet Neurol. 2004, 3, 104–110. [Google Scholar] [CrossRef] [Green Version]
  81. Axisa, P.P.; Hafler, D.A. Multiple sclerosis: Genetics, biomarkers, treatments. Curr. Opin. Neurol. 2016, 29, 345–353. [Google Scholar] [CrossRef] [PubMed]
  82. Baranzini, S.E.; Oksenberg, J.R. The genetics of multiple sclerosis: From 0 to 200 in 50 years. Trends Genet. 2017, 33, 960–970. [Google Scholar] [CrossRef] [PubMed]
  83. Cagliani, R.; Fumagalli, M.; Guerini, F.R.; Riva, S.; Galimberti, D.; Comi, G.P.; Agliardi, C.; Scarpini, E.; Pozzoli, U.; Forni, D.; et al. Identification of a new susceptibility variant for multiple sclerosis in OAS1 by population genetics analysis. Hum. Genet. 2012, 131, 87–97. [Google Scholar] [CrossRef] [PubMed]
  84. Tizaoui, K. Multiple sclerosis genetics: Results from meta-analyses of candidate-gene association studies. Cytokine 2017. [Google Scholar] [CrossRef] [PubMed]
  85. Ryan, J.; Fransquet, P.; Wrigglesworth, J.; Lacaze, P. Phenotypic heterogeneity in dementia: A challenge for epidemiology and biomarker studies. Front. Public Health 2018, 6, 181. [Google Scholar] [CrossRef] [PubMed]
  86. Nelson, P.T.; Head, E.; Schmitt, F.A.; Davis, P.R.; Neltner, J.H.; Jicha, G.A.; Abner, E.L.; Smith, C.D.; Van Eldik, L.J.; Kryscio, R.J. Alzheimer’s disease is not “brain aging”: Neuropathological, genetic, and epidemiological human studies. Acta Neuropathol. 2011, 121, 571–587. [Google Scholar] [CrossRef] [PubMed]
  87. Petrovitch, H.; White, L.; Ross, G.; Steinhorn, S.; Li, C.; Masaki, K.; Davis, D.; Nelson, J.; Hardman, J.; Curb, J. Accuracy of clinical criteria for ad in the Honolulu–Asia aging study, a population-based study. Neurology 2001, 57, 226–234. [Google Scholar] [CrossRef] [PubMed]
  88. McKeith, I.G.; Dickson, D.; Lowe, J.; Emre, M.; O’brien, J.; Feldman, H.; Cummings, J.; Duda, J.; Lippa, C.; Perry, E. Diagnosis and management of dementia with Lewy bodies third report of the DLB consortium. Neurology 2005, 65, 1863–1872. [Google Scholar] [CrossRef] [PubMed]
  89. Delgado-Morales, R.; Esteller, M. Opening up the DNA methylome of dementia. Mol. Psychiatry 2017, 22, 485. [Google Scholar] [CrossRef] [PubMed]
  90. Sperling, R.A.; Karlawish, J.; Johnson, K.A. Preclinical Alzheimer disease—The challenges ahead. Nat. Rev. Neurol. 2013, 9, 54–58. [Google Scholar] [CrossRef] [PubMed]
  91. Citron, M. Alzheimer’s disease: Strategies for disease modification. Nat. Rev. Drug Discov. 2010, 9, 387–398. [Google Scholar] [CrossRef] [PubMed]
  92. Levine, A.J.; Harris, C.R.; Puzio-Kuter, A.M. The interfaces between signal transduction pathways: IGF-1/mTor, p53 and the Parkinson disease pathway. Oncotarget 2012, 3, 1301–1307. [Google Scholar] [CrossRef] [PubMed]
  93. Oddo, S. The role of mTOR signaling in Alzheimer disease. Front. Biosci. 2012, 4, 941–952. [Google Scholar] [CrossRef] [Green Version]
  94. Tramutola, A.; Triplett, J.C.; Di Domenico, F.; Niedowicz, D.M.; Murphy, M.P.; Coccia, R.; Perluigi, M.; Butterfield, D.A. Alteration of mTOR signaling occurs early in the progression of Alzheimer disease (AD): Analysis of brain from subjects with pre-clinical AD, amnestic mild cognitive impairment and late-stage AD. J. Neurochem. 2015, 133, 739–749. [Google Scholar] [CrossRef] [PubMed]
  95. Lan, A.P.; Chen, J.; Zhao, Y.; Chai, Z.; Hu, Y. mTOR signaling in Parkinson’s disease. Neuromol. Med. 2017, 19, 1–10. [Google Scholar] [CrossRef] [PubMed]
  96. Santini, E.; Heiman, M.; Greengard, P.; Valjent, E.; Fisone, G. Inhibition of mtor signaling in Parkinson’s disease prevents L-DOPA-induced dyskinesia. Sci. Signal. 2009, 2, ra36. [Google Scholar] [CrossRef] [PubMed]
  97. Santini, E.; Valjent, E.; Fisone, G. Mtorc1 signaling in Parkinson’s disease and L-DOPA-induced dyskinesia: A sensitized matter. Cell Cycle 2010, 9, 2713–2718. [Google Scholar] [CrossRef] [PubMed]
  98. Cai, Z.; Chen, G.; He, W.; Xiao, M.; Yan, L.J. Activation of mTOR: A culprit of Alzheimer’s disease? Neuropsychiatr. Dis. Treat. 2015, 11, 1015–1030. [Google Scholar] [CrossRef] [PubMed]
  99. Wang, C.; Yu, J.T.; Miao, D.; Wu, Z.C.; Tan, M.S.; Tan, L. Targeting the mTOR signaling network for Alzheimer’s disease therapy. Mol. Neurobiol. 2014, 49, 120–135. [Google Scholar] [CrossRef] [PubMed]
  100. Siman, R.; Cocca, R.; Dong, Y. The mTOR inhibitor rapamycin mitigates perforant pathway neurodegeneration and synapse loss in a mouse model of early-stage Alzheimer-type tauopathy. PLoS ONE 2015, 10, e0142340. [Google Scholar] [CrossRef] [PubMed]
  101. Pryor, W.M.; Biagioli, M.; Shahani, N.; Swarnkar, S.; Huang, W.C.; Page, D.T.; MacDonald, M.E.; Subramaniam, S. Huntingtin promotes mTORc1 signaling in the pathogenesis of Huntington’s disease. Sci. Signal. 2014, 7, ra103. [Google Scholar] [CrossRef] [PubMed]
  102. Ravikumar, B.; Vacher, C.; Berger, Z.; Davies, J.E.; Luo, S.; Oroz, L.G.; Scaravilli, F.; Easton, D.F.; Duden, R.; O’Kane, C.J.; et al. Inhibition of mTOR induces autophagy and reduces toxicity of polyglutamine expansions in fly and mouse models of Huntington disease. Nat. Genet. 2004, 36, 585–595. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Hohlfeld, R.; Wekerle, H. Autoimmune concepts of multiple sclerosis as a basis for selective immunotherapy: From pipe dreams to (therapeutic) pipelines. Proc. Natl. Acad. Sci. USA 2004, 101 (Suppl. 2), 14599–14606. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Frischer, J.M.; Bramow, S.; Dal-Bianco, A.; Lucchinetti, C.F.; Rauschka, H.; Schmidbauer, M.; Laursen, H.; Sorensen, P.S.; Lassmann, H. The relation between inflammation and neurodegeneration in multiple sclerosis brains. Brain J. Neurol. 2009, 132, 1175–1189. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Perl, A. Mtor activation is a biomarker and a central pathway to autoimmune disorders, cancer, obesity, and aging. Ann. N. Y. Acad. Sci. 2015, 1346, 33–44. [Google Scholar] [CrossRef] [PubMed]
  106. Carbone, F.; De Rosa, V.; Carrieri, P.B.; Montella, S.; Bruzzese, D.; Porcellini, A.; Procaccini, C.; La Cava, A.; Matarese, G. Regulatory T cell proliferative potential is impaired in human autoimmune disease. Nat. Med. 2014, 20, 69–74. [Google Scholar] [CrossRef] [PubMed]
  107. Pollizzi, K.N.; Patel, C.H.; Sun, I.H.; Oh, M.H.; Waickman, A.T.; Wen, J.; Delgoffe, G.M.; Powell, J.D. Mtorc1 and mTORc2 selectively regulate CD8(+) T cell differentiation. J. Clin. Investig. 2015, 125, 2090–2108. [Google Scholar] [CrossRef] [PubMed]
  108. Delgoffe, G.M.; Pollizzi, K.N.; Waickman, A.T.; Heikamp, E.; Meyers, D.J.; Horton, M.R.; Xiao, B.; Worley, P.F.; Powell, J.D. The kinase mTOR regulates the differentiation of helper T cells through the selective activation of signaling by mTORc1 and mTORc2. Nat. Immunol. 2011, 12, 295–303. [Google Scholar] [CrossRef] [PubMed]
  109. Tyler, W.A.; Gangoli, N.; Gokina, P.; Kim, H.A.; Covey, M.; Levison, S.W.; Wood, T.L. Activation of the mammalian target of rapamycin (mTOR) is essential for oligodendrocyte differentiation. J. Neurosci. Off. J. Soc. Neurosci. 2009, 29, 6367–6378. [Google Scholar] [CrossRef] [PubMed]
  110. Bercury, K.K.; Dai, J.; Sachs, H.H.; Ahrendsen, J.T.; Wood, T.L.; Macklin, W.B. Conditional ablation of raptor or rictor has differential impact on oligodendrocyte differentiation and cns myelination. J. Neurosci. Off. J. Soc. Neurosci. 2014, 34, 4466–4480. [Google Scholar] [CrossRef] [PubMed]
  111. Lebrun-Julien, F.; Bachmann, L.; Norrmen, C.; Trotzmuller, M.; Kofeler, H.; Ruegg, M.A.; Hall, M.N.; Suter, U. Balanced mTORc1 activity in oligodendrocytes is required for accurate CNS myelination. J. Neurosci. Off. J. Soc. Neurosci. 2014, 34, 8432–8448. [Google Scholar] [CrossRef] [PubMed]
  112. Kim, E.K.; Choi, E.-J. Pathological roles of MAPK signaling pathways in human diseases. Biochim. Biophys. Acta (BBA) Mol. Basis Dis. 2010, 1802, 396–405. [Google Scholar] [CrossRef] [PubMed]
  113. Lengfeld, J.E.; Lutz, S.E.; Smith, J.R.; Diaconu, C.; Scott, C.; Kofman, S.B.; Choi, C.; Walsh, C.M.; Raine, C.S.; Agalliu, I.; et al. Endothelial Wnt/β-catenin signaling reduces immune cell infiltration in multiple sclerosis. Proc. Natl. Acad. Sci. USA 2017, 114, E1168–E1177. [Google Scholar] [CrossRef] [PubMed]
  114. Bayat, V.; Jaiswal, M.; Bellen, H.J. The bmp signaling pathway at the drosophila neuromuscular junction and its links to neurodegenerative diseases. Curr. Opin. Neurobiol. 2011, 21, 182–188. [Google Scholar] [CrossRef] [PubMed]
  115. Holbert, S.; Dedeoglu, A.; Humbert, S.; Saudou, F.; Ferrante, R.J.; Neri, C. Cdc42-interacting protein 4 binds to huntingtin: Neuropathologic and biological evidence for a role in Huntington’s disease. Proc. Natl. Acad. Sci. USA 2003, 100, 2712–2717. [Google Scholar] [CrossRef] [PubMed]
  116. Lione, L.A.; Carter, R.J.; Hunt, M.J.; Bates, G.P.; Morton, A.J.; Dunnett, S.B. Selective discrimination learning impairments in mice expressing the human Huntington’s disease mutation. J. Neurosci. Off. J. Soc. Neurosci. 1999, 19, 10428–10437. [Google Scholar] [CrossRef]
  117. Scheff, S.W.; Price, D.A.; Schmitt, F.A.; Scheff, M.A.; Mufson, E.J. Synaptic loss in the inferior temporal gyrus in mild cognitive impairment and Alzheimer’s disease. J. Alzheimer’s Dis. JAD 2011, 24, 547–557. [Google Scholar] [CrossRef] [PubMed]
  118. Deininger, M.; Meyermann, R.; Schluesener, H. Detection of two transforming growth factor-beta-related morphogens, bone morphogenetic proteins-4 and -5, in RNA of multiple sclerosis and Creutzfeldt-Jakob disease lesions. Acta Neuropathol. 1995, 90, 76–79. [Google Scholar] [CrossRef] [PubMed]
  119. De Ferrari, G.V.; Avila, M.E.; Medina, M.A.; Perez-Palma, E.; Bustos, B.I.; Alarcon, M.A. Wnt/beta-catenin signaling in Alzheimer’s disease. CNS Neurol. Disord. Drug Targets 2014, 13, 745–754. [Google Scholar] [CrossRef] [PubMed]
  120. L’Episcopo, F.; Tirolo, C.; Testa, N.; Caniglia, S.; Morale, M.C.; Serapide, M.F.; Pluchino, S.; Marchetti, B. Wnt/B-catenin signaling is required to rescue midbrain dopaminergic progenitors and promote neurorepair in ageing mouse model of Parkinson’s disease. Stem Cells (Dayton, Ohio) 2014, 32, 2147–2163. [Google Scholar] [CrossRef] [PubMed]
  121. Finkel, S.I. Effects of rivastigmine on behavioral and psychological symptoms of dementia in Alzheimer’s disease. Clin. Ther. 2004, 26, 980–990. [Google Scholar] [CrossRef]
  122. Lee, J.H.; Jeong, S.K.; Kim, B.; Park, K.; Dash, A. Donepezil across the spectrum of Alzheimer’s disease: Dose optimization and clinical relevance. Acta Neurol. Scand. 2015, 131, 259–267. [Google Scholar] [CrossRef] [PubMed]
  123. Quinn, N. Drug treatment of Parkinson’s disease. BMJ Br. Med J. 1995, 310, 575. [Google Scholar] [CrossRef]
  124. Bonuccelli, U.; Colzi, A.; Del Dotto, P. Pergolide in the treatment of patients with early and advanced Parkinson’s disease. Clin. Neuropharmacol. 2002, 25, 1–10. [Google Scholar] [CrossRef] [PubMed]
  125. Le, W.-D.; Jankovic, J. Are dopamine receptor agonists neuroprotective in Parkinson’s disease? Drugs Aging 2001, 18, 389–396. [Google Scholar] [CrossRef] [PubMed]
  126. McMurray, C.T. Huntington’s disease: New hope for therapeutics. Trends Neurosci. 2001, 24, S32–S38. [Google Scholar] [CrossRef]
  127. Faissner, S.; Gold, R. Oral therapies for multiple sclerosis. Cold Spring Harb. Perspect. Med. 2018. [Google Scholar] [CrossRef] [PubMed]
  128. Stangel, M.; Kuhlmann, T.; Matthews, P.M.; Kilpatrick, T.J. Achievements and obstacles of remyelinating therapies in multiple sclerosis. Nat. Rev. Neurol. 2017, 13, 742–754. [Google Scholar] [CrossRef] [PubMed]
  129. Zeydan, B.; Rodriguez, M.; Kantarci, O.H. Timing of future remyelination therapies and their potential to stop multiple sclerosis progression. Adv. Exp. Med. Biol. 2017, 958, 161–170. [Google Scholar] [PubMed]
  130. Plemel, J.R.; Liu, W.Q.; Yong, V.W. Remyelination therapies: A new direction and challenge in multiple sclerosis. Nat. Rev. Drug Discov. 2017, 16, 617–634. [Google Scholar] [CrossRef] [PubMed]
  131. Kumar, V.; Sami, N.; Kashav, T.; Islam, A.; Ahmad, F.; Hassan, M.I. Protein aggregation and neurodegenerative diseases: From theory to therapy. Eur. J. Med. Chem. 2016, 124, 1105–1120. [Google Scholar] [CrossRef] [PubMed]
  132. Herczenik, E.; Gebbink, M.F. Molecular and cellular aspects of protein misfolding and disease. FASEB J. 2008, 22, 2115–2133. [Google Scholar] [CrossRef] [PubMed]
  133. Akiyama, H.; Barger, S.; Barnum, S.; Bradt, B.; Bauer, J.; Cole, G.M.; Cooper, N.R.; Eikelenboom, P.; Emmerling, M.; Fiebich, B.L. Inflammation and Alzheimer’s disease. Neurobiol. Aging 2000, 21, 383–421. [Google Scholar] [CrossRef]
  134. Smith, M.A.; Sayre, L.M.; Monnier, V.M.; Perry, G. Radical ageing in Alzheimer’s disease. Trends Neurosci. 1995, 18, 172–176. [Google Scholar] [CrossRef]
  135. Greenbaum, E.A.; Graves, C.L.; Mishizen-Eberz, A.J.; Lupoli, M.A.; Lynch, D.R.; Englander, S.W.; Axelsen, P.H.; Giasson, B.I. The E46K mutation in α-synuclein increases amyloid fibril formation. J. Boil. Chem. 2005, 280, 7800–7807. [Google Scholar] [CrossRef] [PubMed]
  136. Azuaga, A.I.; Dobson, C.M.; Mateo, P.L.; Conejero-Lara, F. Unfolding and aggregation during the thermal denaturation of streptokinase. FEBS J. 2002, 269, 4121–4133. [Google Scholar] [Green Version]
  137. Iametti, S.; Gregori, B.; Vecchio, G.; Bonomi, F. Modifications occur at different structural levels during the heat denaturation of β-lactoglobulin. FEBS J. 1996, 237, 106–112. [Google Scholar] [CrossRef]
  138. Giri, K.; Bhattacharyya, N.P.; Basak, S. Ph-dependent self-assembly of polyalanine peptides. Biophys. J. 2007, 92, 293–302. [Google Scholar] [CrossRef] [PubMed]
  139. Chaturvedi, S.K.; Alam, P.; Khan, J.M.; Siddiqui, M.K.; Kalaiarasan, P.; Subbarao, N.; Ahmad, Z.; Khan, R.H. Biophysical insight into the anti-amyloidogenic behavior of taurine. Int. J. Boil. Macromol. 2015, 80, 375–384. [Google Scholar] [CrossRef] [PubMed]
  140. Balch, W.E.; Morimoto, R.I.; Dillin, A.; Kelly, J.W. Adapting proteostasis for disease intervention. Science (New York) 2008, 319, 916–919. [Google Scholar] [CrossRef] [PubMed]
  141. Hartl, F.U. Protein misfolding diseases. Annu. Rev. Biochem. 2017, 86, 21–26. [Google Scholar] [CrossRef] [PubMed]
  142. Ross, C.A.; Poirier, M.A. Protein aggregation and neurodegenerative disease. Nat. Med. 2004, 10, S10–S17. [Google Scholar] [CrossRef] [PubMed]
  143. Sherman, M.Y.; Goldberg, A.L. Cellular defenses against unfolded proteins: A cell biologist thinks about neurodegenerative diseases. Neuron 2001, 29, 15–32. [Google Scholar] [CrossRef]
  144. Chen, S.; Brown, I.R. Neuronal expression of constitutive heat shock proteins: Implications for neurodegenerative diseases. Cell Stress Chaperones 2007, 12, 51–58. [Google Scholar] [CrossRef] [PubMed]
  145. Jackrel, M.E.; Shorter, J. Engineering enhanced protein disaggregases for neurodegenerative disease. Prion 2015, 9, 90–109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Hughes, E.; Burke, R.; Doig, A. Inhibition of toxicity in the Alzheimer’s disease peptide fragment β (25–35) using N-methylated derivatives. Biochem. Soc. Trans. 2000, 28, A72. [Google Scholar] [CrossRef]
  147. Madine, J.; Doig, A.J.; Middleton, D.A. Design of an N-methylated peptide inhibitor of α-synuclein aggregation guided by solid-state NMR. J. Am. Chem. Soc. 2008, 130, 7873–7881. [Google Scholar] [CrossRef] [PubMed]
  148. Lu, R.-C.; Tan, M.-S.; Wang, H.; Xie, A.-M.; Yu, J.-T.; Tan, L. Heat shock protein 70 in Alzheimer’s disease. BioMed Res. Int. 2014, 2014, 435203. [Google Scholar] [CrossRef] [PubMed]
  149. Hoshino, T.; Murao, N.; Namba, T.; Takehara, M.; Adachi, H.; Katsuno, M.; Sobue, G.; Matsushima, T.; Suzuki, T.; Mizushima, T. Suppression of Alzheimer’s disease-related phenotypes by expression of heat shock protein 70 in mice. J. Neurosci. 2011, 31, 5225–5234. [Google Scholar] [CrossRef] [PubMed]
  150. Patterson, K.R.; Ward, S.M.; Combs, B.; Voss, K.; Kanaan, N.M.; Morfini, G.; Brady, S.T.; Gamblin, T.C.; Binder, L.I. Heat shock protein 70 prevents both tau aggregation and the inhibitory effects of preexisting tau aggregates on fast axonal transport. Biochemistry 2011, 50, 10300–10310. [Google Scholar] [CrossRef] [PubMed]
  151. Dou, F.; Netzer, W.J.; Tanemura, K.; Li, F.; Hartl, F.U.; Takashima, A.; Gouras, G.K.; Greengard, P.; Xu, H. Chaperones increase association of tau protein with microtubules. Proc. Natl. Acad. Sci. USA 2003, 100, 721–726. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Hirakawa, T.; Rokutan, K.; Nikawa, T.; Kishi, K. Geranylgeranylacetone induces heat shock proteins in cultured guinea pig gastric mucosal cells and rat gastric mucosa. Gastroenterology 1996, 111, 345–357. [Google Scholar] [CrossRef] [PubMed]
  153. Chow, A.M.; Tang, D.W.; Hanif, A.; Brown, I.R. Induction of heat shock proteins in cerebral cortical cultures by celastrol. Cell Stress Chaperones 2013, 18, 155–160. [Google Scholar] [CrossRef] [PubMed]
  154. Bobkova, N.V.; Garbuz, D.G.; Nesterova, I.; Medvinskaya, N.; Samokhin, A.; Alexandrova, I.; Yashin, V.; Karpov, V.; Kukharsky, M.S.; Ninkina, N.N. Therapeutic effect of exogenous hsp70 in mouse models of Alzheimer’s disease. J. Alzheimer’s Dis. 2014, 38, 425–435. [Google Scholar] [CrossRef] [PubMed]
  155. Vinokurov, M.; Ostrov, V.; Yurinskaya, M.; Garbuz, D.; Murashev, A.; Antonova, O.; Evgen’ev, M. Recombinant human hsp70 protects against lipoteichoic acid-induced inflammation manifestations at the cellular and organismal levels. Cell Stress Chaperones 2012, 17, 89–101. [Google Scholar] [CrossRef] [PubMed]
  156. Goldfarb, S.B.; Kashlan, O.B.; Watkins, J.N.; Suaud, L.; Yan, W.; Kleyman, T.R.; Rubenstein, R.C. Differential effects of hsc70 and hsp70 on the intracellular trafficking and functional expression of epithelial sodium channels. Proc. Natl. Acad. Sci. USA 2006, 103, 5817–5822. [Google Scholar] [CrossRef] [PubMed]
  157. Jinwal, U.K.; Akoury, E.; Abisambra, J.F.; O’Leary III, J.C.; Thompson, A.D.; Blair, L.J.; Jin, Y.; Bacon, J.; Nordhues, B.A.; Cockman, M. Imbalance of hsp70 family variants fosters tau accumulation. FASEB J. 2013, 27, 1450–1459. [Google Scholar] [CrossRef] [PubMed]
  158. Miyata, Y.; Li, X.; Lee, H.-F.; Jinwal, U.K.; Srinivasan, S.R.; Seguin, S.P.; Young, Z.T.; Brodsky, J.L.; Dickey, C.A.; Sun, D. Synthesis and initial evaluation of YM-08, a blood-brain barrier permeable derivative of the heat shock protein 70 (Hsp70) inhibitor MKT-077, which reduces tau levels. ACS Chem. Neurosci. 2013, 4, 930–939. [Google Scholar] [CrossRef] [PubMed]
  159. Morimoto, R.I. Dynamic remodeling of transcription complexes by molecular chaperones. Cell 2002, 110, 281–284. [Google Scholar] [CrossRef]
  160. Kim, S.A.; Yoon, J.H.; Lee, S.H.; Ahn, S.G. Polo-like kinase 1 phosphorylates heat shock transcription factor 1 and mediates its nuclear translocation during heat stress. J. Boil. Chem. 2005, 280, 12653–12657. [Google Scholar] [CrossRef] [PubMed]
  161. Hu, Y.; Mivechi, N.F. Hsf-1 interacts with Ral-binding protein 1 in a stress-responsive, multiprotein complex with HSP90 in vivo. J. Boil. Chem. 2003, 278, 17299–17306. [Google Scholar] [CrossRef] [PubMed]
  162. Guo, Y.; Guettouche, T.; Fenna, M.; Boellmann, F.; Pratt, W.B.; Toft, D.O.; Smith, D.F.; Voellmy, R. Evidence for a mechanism of repression of heat shock factor 1 transcriptional activity by a multichaperone complex. J. Boil. Chem. 2001, 276, 45791–45799. [Google Scholar] [CrossRef] [PubMed]
  163. Cushman, M.; Johnson, B.S.; King, O.D.; Gitler, A.D.; Shorter, J. Prion-like disorders: Blurring the divide between transmissibility and infectivity. J. Cell Sci. 2010, 123, 1191–1201. [Google Scholar] [CrossRef] [PubMed]
  164. Lindberg, I.; Shorter, J.; Wiseman, R.L.; Chiti, F.; Dickey, C.A.; McLean, P.J. Chaperones in neurodegeneration. J. Neurosci. 2015, 35, 13853–13859. [Google Scholar] [CrossRef] [PubMed]
  165. Takalo, M.; Salminen, A.; Soininen, H.; Hiltunen, M.; Haapasalo, A. Protein aggregation and degradation mechanisms in neurodegenerative diseases. Am. J. Neurodegener. Dis. 2013, 2, 1–14. [Google Scholar] [PubMed]
  166. Shorter, J. Engineering therapeutic protein disaggregases. Mol. Boil. Cell 2016, 27, 1556–1560. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Shorter, J. Hsp104: A weapon to combat diverse neurodegenerative disorders. Neurosignals 2008, 16, 63–74. [Google Scholar] [CrossRef] [PubMed]
  168. Jackrel, M.E.; Shorter, J. Potentiated Hsp104 variants suppress toxicity of diverse neurodegenerative disease-linked proteins. Dis. Model Mech. 2014, 7, 1175–1184. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Lee, S.J.; Lim, H.S.; Masliah, E.; Lee, H.J. Protein aggregate spreading in neurodegenerative diseases: Problems and perspectives. Neurosci. Res. 2011, 70, 339–348. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Evans, C.G.; Wisén, S.; Gestwicki, J.E. Heat shock proteins 70 and 90 inhibit early stages of amyloid beta (1-42) aggregation in vitro. J. Boil. Chem. 2006. [Google Scholar] [CrossRef] [PubMed]
  171. Bauer, P.O.; Goswami, A.; Wong, H.K.; Okuno, M.; Kurosawa, M.; Yamada, M.; Miyazaki, H.; Matsumoto, G.; Kino, Y.; Nagai, Y.; et al. Harnessing chaperone-mediated autophagy for the selective degradation of mutant huntingtin protein. Nat. Biotechnol. 2010, 28, 256–263. [Google Scholar] [CrossRef] [PubMed]
  172. Koyuncu, S.; Fatima, A.; Gutierrez-Garcia, R.; Vilchez, D. Proteostasis of huntingtin in health and disease. Int. J. Mol. Sci. 2017, 18, 1568. [Google Scholar] [CrossRef] [PubMed]
  173. Maiti, P.; Manna, J.; Veleri, S.; Frautschy, S. Molecular chaperone dysfunction in neurodegenerative diseases and effects of curcumin. BioMed Res. Int. 2014, 2014, 495091. [Google Scholar] [CrossRef] [PubMed]
  174. Dedmon, M.M.; Christodoulou, J.; Wilson, M.R.; Dobson, C.M. Heat shock protein 70 inhibits alpha-synuclein fibril formation via preferential binding to prefibrillar species. J. Boil. Chem. 2005, 280, 14733–14740. [Google Scholar] [CrossRef] [PubMed]
  175. Ebrahimi-Fakhari, D.; Wahlster, L.; McLean, P.J. Molecular chaperones in Parkinson’s disease—Present and future. J. Park. Dis. 2011, 1, 299–320. [Google Scholar]
  176. Schulte, T.W.; Neckers, L.M. The benzoquinone ansamycin 17-allylamino-17-demethoxygeldanamycin binds to HSP90 and shares important biologic activities with geldanamycin. Cancer Chemother. Pharmacol. 1998, 42, 273–279. [Google Scholar] [CrossRef] [PubMed]
  177. Shen, H.Y.; He, J.C.; Wang, Y.; Huang, Q.Y.; Chen, J.F. Geldanamycin induces heat shock protein 70 and protects against MPTP-induced dopaminergic neurotoxicity in mice. J. Boil. Chem. 2005, 280, 39962–39969. [Google Scholar] [CrossRef] [PubMed]
  178. Ciechanover, A.; Kwon, Y.T. Protein quality control by molecular chaperones in neurodegeneration. Front. Neurosci. 2017, 11, 185. [Google Scholar] [CrossRef] [PubMed]
  179. Björkqvist, M.; Wild, E.J.; Thiele, J.; Silvestroni, A.; Andre, R.; Lahiri, N.; Raibon, E.; Lee, R.V.; Benn, C.L.; Soulet, D.; et al. A novel pathogenic pathway of immune activation detectable before clinical onset in Huntington’s disease. J. Exp. Med. 2008, 205, 1869–1877. [Google Scholar] [CrossRef] [PubMed]
  180. Chen, H.; O’Reilly, E.J.; Schwarzschild, M.A.; Ascherio, A. Peripheral inflammatory biomarkers and risk of Parkinson’s disease. Am. J. Epidemiol. 2008, 167, 90–95. [Google Scholar] [CrossRef] [PubMed]
  181. Mogi, M.; Harada, M.; Riederer, P.; Narabayashi, H.; Fujita, K.; Nagatsu, T. Tumor necrosis factor-alpha (TNF-alpha) increases both in the brain and in the cerebrospinal fluid from parkinsonian patients. Neurosci. Lett. 1994, 165, 208–210. [Google Scholar] [CrossRef]
  182. Crotti, A.; Glass, C.K. The choreography of neuroinflammation in Huntington’s disease. Trends Immunol. 2015, 36, 364–373. [Google Scholar] [CrossRef] [PubMed]
  183. Tronel, C.; Largeau, B.; Santiago Ribeiro, M.J.; Guilloteau, D.; Dupont, A.-C.; Arlicot, N. Molecular targets for pet imaging of activated microglia: The current situation and future expectations. Int. J. Mol. Sci. 2017, 18, 802. [Google Scholar] [CrossRef] [PubMed]
  184. Tai, Y.F.; Pavese, N.; Gerhard, A.; Tabrizi, S.J.; Barker, R.A.; Brooks, D.J.; Piccini, P. Microglial activation in presymptomatic Huntington’s disease gene carriers. Brain J. Neurol. 2007, 130, 1759–1766. [Google Scholar] [CrossRef] [PubMed]
  185. Rocha, N.P.; Ribeiro, F.M.; Furr-Stimming, E.; Teixeira, A.L. Neuroimmunology of Huntington’s disease: Revisiting evidence from human studies. Mediat. Inflamm. 2016, 2016, 8653132. [Google Scholar] [CrossRef] [PubMed]
  186. Sapp, E.; Kegel, K.B.; Aronin, N.; Hashikawa, T.; Uchiyama, Y.; Tohyama, K.; Bhide, P.G.; Vonsattel, J.P.; DiFiglia, M. Early and progressive accumulation of reactive microglia in the Huntington disease brain. J. Neuropathol. Exp. Neurol. 2001, 60, 161–172. [Google Scholar] [CrossRef] [PubMed]
  187. Miron, V.E.; Boyd, A.; Zhao, J.W.; Yuen, T.J.; Ruckh, J.M.; Shadrach, J.L.; van Wijngaarden, P.; Wagers, A.J.; Williams, A.; Franklin, R.J.M.; et al. M2 microglia and macrophages drive oligodendrocyte differentiation during CNS remyelination. Nat. Neurosci. 2013, 16, 1211–1218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Yang, H.-M.; Yang, S.; Huang, S.-S.; Tang, B.-S.; Guo, J.-F. Microglial activation in the pathogenesis of Huntington’s disease. Front. Aging Neurosci. 2017, 9, 193. [Google Scholar] [CrossRef] [PubMed]
  189. Chang, K.H.; Wu, Y.R.; Chen, Y.C.; Chen, C.M. Plasma inflammatory biomarkers for Huntington’s disease patients and mouse model. Brain Behav. Immun. 2015, 44, 121–127. [Google Scholar] [CrossRef] [PubMed]
  190. Politis, M.; Lahiri, N.; Niccolini, F.; Su, P.; Wu, K.; Giannetti, P.; Scahill, R.I.; Turkheimer, F.E.; Tabrizi, S.J.; Piccini, P. Increased central microglial activation associated with peripheral cytokine levels in premanifest Huntington’s disease gene carriers. Neurobiol. Dis. 2015, 83, 115–121. [Google Scholar] [CrossRef] [PubMed]
  191. Träger, U.; Andre, R.; Lahiri, N.; Magnusson-Lind, A.; Weiss, A.; Grueninger, S.; McKinnon, C.; Sirinathsinghji, E.; Kahlon, S.; Pfister, E.L. HTT-lowering reverses Huntington’s disease immune dysfunction caused by NFκB pathway dysregulation. Brain J. Neurol. 2014, 137, 819–833. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Brochard, V.; Combadiere, B.; Prigent, A.; Laouar, Y.; Perrin, A.; Beray-Berthat, V.; Bonduelle, O.; Alvarez-Fischer, D.; Callebert, J.; Launay, J.M.; et al. Infiltration of CD4+ lymphocytes into the brain contributes to neurodegeneration in a mouse model of Parkinson disease. J. Clin. Investig. 2009, 119, 182–192. [Google Scholar] [CrossRef] [PubMed]
  193. Rees, K.; Stowe, R.; Patel, S.; Ives, N.; Breen, K.; Clarke, C.E.; Ben-Shlomo, Y. Non-steroidal anti-inflammatory drugs as disease-modifying agents for Parkinson’s disease: Evidence from observational studies. Cochrane Database Syst. Rev. 2011, Cd008454. [Google Scholar]
  194. Du, Y.; Ma, Z.; Lin, S.; Dodel, R.C.; Gao, F.; Bales, K.R.; Triarhou, L.C.; Chernet, E.; Perry, K.W.; Nelson, D.L.; et al. Minocycline prevents nigrostriatal dopaminergic neurodegeneration in the MPTP model of Parkinson’s disease. Proc. Natl. Acad. Sci. USA 2001, 98, 14669–14674. [Google Scholar] [CrossRef] [PubMed]
  195. Lofrumento, D.D.; Nicolardi, G.; Cianciulli, A.; Nuccio, F.D.; Pesa, V.L.; Carofiglio, V.; Dragone, T.; Calvello, R.; Panaro, M.A. Neuroprotective effects of resveratrol in an MPTP mouse model of Parkinson’s-like disease: Possible role of socs-1 in reducing pro-inflammatory responses. Innate Immun. 2014, 20, 249–260. [Google Scholar] [CrossRef] [PubMed]
  196. Zhang, F.; Shi, J.-S.; Zhou, H.; Wilson, B.C.; Hong, J.-S.; Gao, H.-M. Resveratrol protects dopamine neurons against lipopolysaccharide-induced neurotoxicity through its anti-inflammatory actions. Mol. Pharmacol. 2010, 94. [Google Scholar] [CrossRef] [PubMed]
  197. Ren, B.; Zhang, Y.-X.; Zhou, H.-X.; Sun, F.-W.; Zhang, Z.-F.; Wei, Z.-F.; Zhang, C.-Y.; Si, D.-W. Tanshinone IIA prevents the loss of nigrostriatal dopaminergic neurons by inhibiting NADPH oxidase and iNOS in the MPTP model of Parkinson’s disease. J. Neurol. Sci. 2015, 348, 142–152. [Google Scholar] [CrossRef] [PubMed]
  198. Wang, S.; Jing, H.; Yang, H.; Liu, Z.; Guo, H.; Chai, L.; Hu, L. Tanshinone I selectively suppresses pro-inflammatory genes expression in activated microglia and prevents nigrostriatal dopaminergic neurodegeneration in a mouse model of Parkinson’s disease. J. Ethnopharmacol. 2015, 164, 247–255. [Google Scholar] [CrossRef] [PubMed]
  199. Borah, A.; Paul, R.; Choudhury, S.; Choudhury, A.; Bhuyan, B.; Das Talukdar, A.; Dutta Choudhury, M.; Mohanakumar, K.P. Neuroprotective potential of silymarin against CNS disorders: Insight into the pathways and molecular mechanisms of action. CNS Neurosci. Ther. 2013, 19, 847–853. [Google Scholar] [CrossRef] [PubMed]
  200. Gendelman, H.E.; Zhang, Y.; Santamaria, P.; Olson, K.E.; Schutt, C.R.; Bhatti, D.; Shetty, B.L.D.; Lu, Y.; Estes, K.A.; Standaert, D.G. Evaluation of the safety and immunomodulatory effects of sargramostim in a randomized, double-blind phase 1 clinical Parkinson’s disease trial. NPJ Park. Dis. 2017, 3, 10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  201. Dobson, L.; Träger, U.; Farmer, R.; Hayardeny, L.; Loupe, P.; Hayden, M.R.; Tabrizi, S.J. Laquinimod dampens hyperactive cytokine production in Huntington’s disease patient myeloid cells. J. Neurochem. 2016, 137, 782–794. [Google Scholar] [CrossRef] [PubMed]
  202. Na, S.Y.; Mracsko, E.; Liesz, A.; Hunig, T.; Veltkamp, R. Amplification of regulatory t cells using a CD28 superagonist reduces brain damage after ischemic stroke in mice. Stroke 2015, 46, 212–220. [Google Scholar] [CrossRef] [PubMed]
  203. Chung, E.S.; Lee, G.; Lee, C.; Ye, M.; Chung, H.S.; Kim, H.; Bae, S.J.; Hwang, D.S.; Bae, H. Bee venom phospholipase A2, a novel Foxp3+ regulatory t cell inducer, protects dopaminergic neurons by modulating neuroinflammatory responses in a mouse model of Parkinson’s disease. J. Immunol. 2015, 195, 4853–4860. [Google Scholar] [CrossRef] [PubMed]
  204. Games, D.; Valera, E.; Spencer, B.; Rockenstein, E.; Mante, M.; Adame, A.; Patrick, C.; Ubhi, K.; Nuber, S.; Sacayon, P. Reducing C-terminal-truncated alpha-synuclein by immunotherapy attenuates neurodegeneration and propagation in Parkinson’s disease-like models. J. Neurosci. 2014, 34, 9441–9454. [Google Scholar] [CrossRef] [PubMed]
  205. Lindström, V.; Fagerqvist, T.; Nordström, E.; Eriksson, F.; Lord, A.; Tucker, S.; Andersson, J.; Johannesson, M.; Schell, H.; Kahle, P.J. Immunotherapy targeting α-synuclein protofibrils reduced pathology in (Thy-1)-h [A30P] α-synuclein mice. Neurobiol. Dis. 2014, 69, 134–143. [Google Scholar] [CrossRef] [PubMed]
  206. O’Nuallain, B.; Williams, A.D.; McWilliams-Koeppen, H.P.; Acero, L.; Weber, A.; Ehrlich, H.; Schwarz, H.P.; Solomon, A. Anti-amyloidogenic activity of IgGs contained in normal plasma. J. Clin. Immunol. 2010, 30, 37–42. [Google Scholar] [CrossRef] [PubMed]
  207. Kayed, R.; Canto, I.; Breydo, L.; Rasool, S.; Lukacsovich, T.; Wu, J.; Albay, R.; Pensalfini, A.; Yeung, S.; Head, E. Conformation dependent monoclonal antibodies distinguish different replicating strains or conformers of prefibrillar Aβ oligomers. Mol. Neurodegener. 2010, 5, 57. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Lambert, M.P.; Velasco, P.T.; Chang, L.; Viola, K.L.; Fernandez, S.; Lacor, P.N.; Khuon, D.; Gong, Y.; Bigio, E.H.; Shaw, P. Monoclonal antibodies that target pathological assemblies of Aβ. J. Neurochem. 2007, 100, 23–35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Mizushima, N.; Levine, B.; Cuervo, A.M.; Klionsky, D.J. Autophagy fights disease through cellular self-digestion. Nature 2008, 451, 1069–1075. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Yen, W.-L.; Klionsky, D.J. How to live long and prosper: Autophagy, mitochondria, and aging. Physiology 2008, 23, 248–262. [Google Scholar] [CrossRef] [PubMed]
  211. Boland, B.; Kumar, A.; Lee, S.; Platt, F.M.; Wegiel, J.; Yu, W.H.; Nixon, R.A. Autophagy induction and autophagosome clearance in neurons: Relationship to autophagic pathology in Alzheimer’s disease. J. Neurosci. 2008, 28, 6926–6937. [Google Scholar] [CrossRef] [PubMed]
  212. Nixon, R.A. The role of autophagy in neurodegenerative disease. Nat. Med. 2013, 19, 983–997. [Google Scholar] [CrossRef] [PubMed]
  213. Stoica, L.; Zhu, P.J.; Huang, W.; Zhou, H.; Kozma, S.C.; Costa-Mattioli, M. Selective pharmacogenetic inhibition of mammalian target of rapamycin complex I (mTORc1) blocks long-term synaptic plasticity and memory storage. Proc. Natl. Acad. Sci. USA 2011, 108, 3791–3796. [Google Scholar] [CrossRef] [PubMed]
  214. Pupyshev, A.B.; Korolenko, T.A.; Tikhonova, M.A. A therapeutic target for inhibition of neurodegeneration: Autophagy. Neurosci. Behav. Physiol. 2017, 47, 1109–1127. [Google Scholar] [CrossRef]
  215. Nah, J.; Yuan, J.; Jung, Y.-K. Autophagy in neurodegenerative diseases: From mechanism to therapeutic approach. Mol. Cells 2015, 38, 381–389. [Google Scholar] [CrossRef] [PubMed]
  216. Hara, T.; Nakamura, K.; Matsui, M.; Yamamoto, A.; Nakahara, Y.; Suzuki-Migishima, R.; Yokoyama, M.; Mishima, K.; Saito, I.; Okano, H. Suppression of basal autophagy in neural cells causes neurodegenerative disease in mice. Nature 2006, 441, 885–889. [Google Scholar] [CrossRef] [PubMed]
  217. Liang, C.-C.; Wang, C.; Peng, X.; Gan, B.; Guan, J.-L. Neural-specific deletion of FIP200 leads to cerebellar degeneration caused by increased neuronal death and axon degeneration. J. Boil. Chem. 2010, 285, 3499–3509. [Google Scholar] [CrossRef] [PubMed]
  218. Choi, S.J.; Panhelainen, A.; Schmitz, Y.; Larsen, K.E.; Kanter, E.; Wu, M.; Sulzer, D.; Mosharov, E.V. Changes in neuronal dopamine homeostasis following 1-methyl-4-phenylpyridinium (MPP+) exposure. J. Boil. Chem. 2015, 290, 6799–6809. [Google Scholar] [CrossRef] [PubMed]
  219. Javitch, J.A.; D’Amato, R.J.; Strittmatter, S.M.; Snyder, S.H. Parkinsonism-inducing neurotoxin, N-methyl-4-phenyl-1,2,3,6-tetrahydropyridine: Uptake of the metabolite N-methyl-4-phenylpyridine by dopamine neurons explains selective toxicity. Proc. Natl. Acad. Sci. USA 1985, 82, 2173–2177. [Google Scholar] [CrossRef] [PubMed]
  220. Steele, J.W.; Gandy, S. Latrepirdine (Dimebon(R)), a potential alzheimer therapeutic, regulates autophagy and neuropathology in an Alzheimer mouse model. Autophagy 2013, 9, 617–618. [Google Scholar] [CrossRef] [PubMed]
  221. Chau, S.; Herrmann, N.; Ruthirakuhan, M.T.; Chen, J.J.; Lanctot, K.L. Latrepirdine for Alzheimer’s Disease; The Cochrane Library: London, UK, 2015. [Google Scholar]
  222. Forlenza, O.V.; de Paula, V.J.; Machado-Vieira, R.; Diniz, B.S.; Gattaz, W.F. Does lithium prevent Alzheimer’s disease? Drugs Aging 2012, 29, 335–342. [Google Scholar] [CrossRef] [PubMed]
  223. Berger, Z.; Ravikumar, B.; Menzies, F.M.; Oroz, L.G.; Underwood, B.R.; Pangalos, M.N.; Schmitt, I.; Wullner, U.; Evert, B.O.; O’kane, C.J. Rapamycin alleviates toxicity of different aggregate-prone proteins. Hum. Mol. Genet. 2005, 15, 433–442. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Wang, T.; Lao, U.; Edgar, B.A. Tor-mediated autophagy regulates cell death in drosophila neurodegenerative disease. J. Cell Boil. 2009, 186, 703–711. [Google Scholar] [CrossRef] [PubMed]
  225. Park, H.-W.; Park, H.; Semple, I.A.; Jang, I.; Ro, S.-H.; Kim, M.; Cazares, V.A.; Stuenkel, E.L.; Kim, J.-J.; Kim, J.S. Pharmacological correction of obesity-induced autophagy arrest using calcium channel blockers. Nat. Commun. 2014, 5, 4834. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Caccamo, A.; Majumder, S.; Richardson, A.; Strong, R.; Oddo, S. Molecular interplay between mammalian target of rapamycin (mTOR), amyloid-beta, and tau: Effects on cognitive impairments. J. Boil. Chem. 2010, 285, 13107–13120. [Google Scholar] [CrossRef] [PubMed]
  227. Sun, Q.; Wei, L.-L.; Zhang, M.; Li, T.-X.; Yang, C.; Deng, S.-P.; Zeng, Q.-C. Rapamycin inhibits activation of ampk-mTOR signaling pathway-induced Alzheimer’s disease lesion in hippocampus of rats with type 2 diabetes mellitus. Int. J. Neurosci. 2018, 2018, 1–22. [Google Scholar] [CrossRef] [PubMed]
  228. Wullschleger, S.; Loewith, R.; Hall, M.N. Tor signaling in growth and metabolism. Cell 2006, 124, 471–484. [Google Scholar] [CrossRef] [PubMed]
  229. Ma, T.C.; Buescher, J.L.; Oatis, B.; Funk, J.A.; Nash, A.J.; Carrier, R.L.; Hoyt, K.R. Metformin therapy in a transgenic mouse model of Huntington’s disease. Neurosci. Lett. 2007, 411, 98–103. [Google Scholar]
  230. Goedert, M.; Jakes, R.; Qi, Z.; Wang, J.H.; Cohen, P. Protein phosphatase 2A is the major enzyme in brain that dephosphorylates tau protein phosphorylated by proline-directed protein kinases or cyclic amp-dependent protein kinase. J Neurochem. 1995, 65, 2804–2807. [Google Scholar] [CrossRef] [PubMed]
  231. Hebron, M.L.; Lonskaya, I.; Olopade, P.; Selby, S.T.; Pagan, F.; Moussa, C.E. Tyrosine kinase inhibition regulates early systemic immune changes and modulates the neuroimmune response in α-synucleinopathy. J. Clin. Cell. Immunol. 2014, 5, 259. [Google Scholar] [CrossRef] [PubMed]
  232. Lonskaya, I.; Hebron, M.; Selby, S.; Turner, R.; Moussa, C.-H. Nilotinib and Bosutinib modulate pre-plaque alterations of blood immune markers and neuro-inflammation in Alzheimer’s disease models. Neuroscience 2015, 304, 316–327. [Google Scholar] [CrossRef] [PubMed]
  233. Iliff, J.J.; Nedergaard, M. Is there a cerebral lymphatic system? Stroke 2013, 44, S93–S95. [Google Scholar] [CrossRef] [PubMed]
  234. Xie, L.; Kang, H.; Xu, Q.; Chen, M.J.; Liao, Y.; Thiyagarajan, M.; O’Donnell, J.; Christensen, D.J.; Nicholson, C.; Iliff, J.J.; et al. Sleep drives metabolite clearance from the adult brain. Science (New York) 2013, 342, 373–377. [Google Scholar] [CrossRef] [PubMed]
  235. Nedergaard, M. Neuroscience. Garbage truck of the brain. Science (New York) 2013, 340, 1529–1530. [Google Scholar] [CrossRef] [PubMed]
  236. Iliff, J.J.; Lee, H.; Yu, M.; Feng, T.; Logan, J.; Nedergaard, M.; Benveniste, H. Brain-wide pathway for waste clearance captured by contrast-enhanced MRI. J. Clin. Investig. 2013, 123, 1299–1309. [Google Scholar] [CrossRef] [PubMed]
  237. Jessen, N.A.; Munk, A.S.; Lundgaard, I.; Nedergaard, M. The glymphatic system: A beginner’s guide. Neurochem. Res. 2015, 40, 2583–2599. [Google Scholar] [CrossRef] [PubMed]
  238. Kress, B.T.; Iliff, J.J.; Xia, M.; Wang, M.; Wei, H.S.; Zeppenfeld, D.; Xie, L.; Kang, H.; Xu, Q.; Liew, J.A.; et al. Impairment of paravascular clearance pathways in the aging brain. Ann. Neurol. 2014, 76, 845–861. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  239. Iliff, J.J.; Wang, M.; Liao, Y.; Plogg, B.A.; Peng, W.; Gundersen, G.A.; Benveniste, H.; Vates, G.E.; Deane, R.; Goldman, S.A.; et al. A paravascular pathway facilitates CSF flow through the brain parenchyma and the clearance of interstitial solutes, including amyloid beta. Sci. Transl. Med. 2012, 4, 147ra111. [Google Scholar] [CrossRef] [PubMed]
  240. Nedergaard, M.; Goldman, S.A. Brain drain. Sci. Am. 2016, 314, 44–49. [Google Scholar] [CrossRef] [PubMed]
  241. de Leon, M.J.; Li, Y.; Okamura, N.; Tsui, W.H.; Saint Louis, L.A.; Glodzik, L.; Osorio, R.S.; Fortea, J.; Butler, T.; Pirraglia, E.; et al. CSF clearance in Alzheimer disease measured with dynamic pet. J. Nucl. Med. 2017, 58, 1471. [Google Scholar] [CrossRef] [PubMed]
  242. Peng, W.; Achariyar, T.M.; Li, B.; Liao, Y.; Mestre, H.; Hitomi, E.; Regan, S.; Kasper, T.; Peng, S.; Ding, F.; et al. Suppression of glymphatic fluid transport in a mouse model of Alzheimer’s disease. Neurobiol. Dis. 2016, 93, 215–225. [Google Scholar] [CrossRef] [PubMed]
  243. Oh, H.; Madison, C.; Baker, S.; Rabinovici, G.; Jagust, W. Dynamic relationships between age, amyloid-beta deposition, and glucose metabolism link to the regional vulnerability to Alzheimer’s disease. Brain J. Neurol. 2016, 139, 2275–2289. [Google Scholar] [CrossRef] [PubMed]
  244. Gage, F.H. Neurogenesis in the adult brain. J. Neurosci. 2002, 22, 612–613. [Google Scholar] [CrossRef] [PubMed]
  245. Reynolds, B.A.; Weiss, S. Generation of neurons and astrocytes from isolated cells of the adult mammalian central nervous system. Science (New York) 1992, 255, 1707–1710. [Google Scholar] [CrossRef]
  246. Shihabuddin, L.S.; Horner, P.J.; Ray, J.; Gage, F.H. Adult spinal cord stem cells generate neurons after transplantation in the adult dentate gyrus. J. Neurosci. Off. J. Soc. Neurosci. 2000, 20, 8727–8735. [Google Scholar] [CrossRef]
  247. Pencea, V.; Bingaman, K.D.; Freedman, L.J.; Luskin, M.B. Neurogenesis in the subventricular zone and rostral migratory stream of the neonatal and adult primate forebrain. Exp. Neurol. 2001, 172, 1–16. [Google Scholar] [CrossRef] [PubMed]
  248. Allen, S.J.; Watson, J.J.; Dawbarn, D. The neurotrophins and their role in Alzheimer’s disease. Curr. Neuropharmacol. 2011, 9, 559–573. [Google Scholar] [CrossRef] [PubMed]
  249. Lindsay, R.M.; Altar, C.A.; Cedarbaum, J.M.; Hyman, C.; Wiegand, S.J. The therapeutic potential of neurotrophic factors in the treatment of Parkinson’s disease. Exp. Neurol. 1993, 124, 103–118. [Google Scholar] [CrossRef] [PubMed]
  250. Grondin, R.; Gash, D.M. Glial cell line-derived neurotrophic factor (GDNF): A drug candidate for the treatment of Parkinson’s disease. J. Neurol. 1998, 245, P35–P42. [Google Scholar] [CrossRef] [PubMed]
  251. Bowenkamp, K.E.; Lapchak, P.A.; Hoffer, B.J.; Miller, P.J.; Bickford, P.C. Intracerebroventricular glial cell line-derived neurotrophic factor improves motor function and supports nigrostriatal dopamine neurons in bilaterally 6-hydroxydopamine lesioned rats. Exp. Neurol. 1997, 145, 104–117. [Google Scholar] [CrossRef] [PubMed]
  252. Duman, R.S. Neuronal damage and protection in the pathophysiology and treatment of psychiatric illness: Stress and depression. Dialogues Clin. Neurosci. 2009, 11, 239–255. [Google Scholar] [PubMed]
  253. Tsiperson, V.; Huang, Y.; Bagayogo, I.; Song, Y.; VonDran, M.W.; DiCicco-Bloom, E.; Dreyfus, C.F. Brain-derived neurotrophic factor deficiency restricts proliferation of oligodendrocyte progenitors following cuprizone-induced demyelination. ASN Neuro 2015, 7. [Google Scholar] [CrossRef] [PubMed]
  254. Schindowski, K.; Belarbi, K.; Buée, L. Neurotrophic factors in Alzheimer’s disease: Role of axonal transport. Genes Brain Behav. 2008, 7, 43–56. [Google Scholar] [CrossRef] [PubMed]
  255. Jiao, S.S.; Shen, L.L.; Zhu, C.; Bu, X.L.; Liu, Y.H.; Liu, C.H.; Yao, X.Q.; Zhang, L.L.; Zhou, H.D.; Walker, D.G.; et al. Brain-derived neurotrophic factor protects against tau-related neurodegeneration of Alzheimer’s disease. Transl. Psychiatry 2016, 6, e907. [Google Scholar] [CrossRef] [PubMed]
  256. Gold, S.M.; Chalifoux, S.; Giesser, B.S.; Voskuhl, R.R. Immune modulation and increased neurotrophic factor production in multiple sclerosis patients treated with testosterone. J. Neuroinflamm. 2008, 5, 32. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Cohen-Cory, S.; Kidane, A.H.; Shirkey, N.J.; Marshak, S. Brain-derived neurotrophic factor and the development of structural neuronal connectivity. Dev. Neurobiol. 2010, 70, 271–288. [Google Scholar] [CrossRef] [PubMed]
  258. Henderson, C.E. Role of neurotrophic factors in neuronal development. Curr. Opin. Neurobiol. 1996, 6, 64–70. [Google Scholar] [CrossRef]
  259. Poo, M.M. Neurotrophins as synaptic modulators. Nat. Rev. Neurosci. 2001, 2, 24–32. [Google Scholar] [CrossRef] [PubMed]
  260. Huang, E.J.; Reichardt, L.F. Neurotrophins: Roles in neuronal development and function. Annu. Rev. Neurosci. 2001, 24, 677–736. [Google Scholar] [CrossRef] [PubMed]
  261. Nguyen, T.L.X.; Kim, C.K.; Cho, J.-H.; Lee, K.-H.; Ahn, J.-Y. Neuroprotection signaling pathway of nerve growth factor and brain-derived neurotrophic factor against staurosporine induced apoptosis in hippocampal H19-7 cells. Exp. Mol. Med. 2010, 42, 583–595. [Google Scholar] [CrossRef] [PubMed]
  262. Middlemas, D.S.; Kihl, B.K.; Zhou, J.; Zhu, X. Brain-derived neurotrophic factor promotes survival and chemoprotection of human neuroblastoma cells. J. Boil. Chem. 1999, 274, 16451–16460. [Google Scholar] [CrossRef]
  263. Weissmiller, A.M.; Wu, C. Current advances in using neurotrophic factors to treat neurodegenerative disorders. Transl. Neurodegener. 2012, 1, 14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Salehi, A.; Delcroix, J.D.; Swaab, D.F. Alzheimer’s disease and NGF signaling. J. Neural Transm. (Vienna, Austria 1996) 2004, 111, 323–345. [Google Scholar] [CrossRef] [PubMed]
  265. Higgins, G.A.; Mufson, E.J. Ngf receptor gene expression is decreased in the nucleus basalis in Alzheimer’s disease. Exp. Neurol. 1989, 106, 222–236. [Google Scholar] [CrossRef]
  266. Costa, A.; Peppe, A.; Carlesimo, G.A.; Zabberoni, S.; Scalici, F.; Caltagirone, C.; Angelucci, F. Brain-derived neurotrophic factor serum levels correlate with cognitive performance in Parkinson’s disease patients with mild cognitive impairment. Front. Behav. Neurosci. 2015, 9, 253. [Google Scholar] [CrossRef] [PubMed]
  267. Howells, D.W.; Porritt, M.J.; Wong, J.Y.; Batchelor, P.E.; Kalnins, R.; Hughes, A.J.; Donnan, G.A. Reduced bdnf mrna expression in the Parkinson’s disease substantia nigra. Exp. Neurol. 2000, 166, 127–135. [Google Scholar] [CrossRef] [PubMed]
  268. Pan, W.; Banks, W.A.; Kastin, A.J. Permeability of the blood–brain barrier to neurotrophins. Brain Res. 1998, 788, 87–94. [Google Scholar] [CrossRef]
  269. Pollack, S.J.; Harper, S.J. Small molecule Trk receptor agonists and other neurotrophic factor mimetics. Curr. Drug Targets CNS Neurol. Disord. 2002, 1, 59–80. [Google Scholar] [CrossRef] [PubMed]
  270. Baulieu, E.E.; Robel, P. Neurosteroids: A new brain function? J. Steroid Biochem. Mol. Boil. 1990, 37, 395–403. [Google Scholar] [CrossRef]
  271. Lv, W.; Du, N.; Liu, Y.; Fan, X.; Wang, Y.; Jia, X.; Hou, X.; Wang, B. Low testosterone level and risk of Alzheimer’s disease in the elderly men: A systematic review and meta-analysis. Mol. Neurobiol. 2016, 53, 2679–2684. [Google Scholar] [CrossRef] [PubMed]
  272. Paganini-Hill, A.; Henderson, V.W. Estrogen deficiency and risk of Alzheimer’s disease in women. Am. J. Epidemiol. 1994, 140, 256–261. [Google Scholar] [CrossRef] [PubMed]
  273. Carroll, J.C.; Rosario, E.R.; Villamagna, A.; Pike, C.J. Continuous and cyclic progesterone differentially interact with estradiol in the regulation of Alzheimer-like pathology in female 3xtransgenic-Alzheimer’s disease mice. Endocrinology 2010, 151, 2713–2722. [Google Scholar] [CrossRef] [PubMed]
  274. Rosario, E.R.; Chang, L.; Stanczyk, F.Z.; Pike, C.J. Age-related testosterone depletion and the development of Alzheimer disease. JAMA 2004, 292, 1431–1432. [Google Scholar] [CrossRef] [PubMed]
  275. Seidl, J.N.; Massman, P.J. Relationships between testosterone levels and cognition in patients with Alzheimer disease and nondemented elderly men. J. Geriatr. Psychiatry Neurol. 2015, 28, 27–39. [Google Scholar] [CrossRef] [PubMed]
  276. Okun, M.S.; DeLong, M.R.; Hanfelt, J.; Gearing, M.; Levey, A. Plasma testosterone levels in Alzheimer and Parkinson diseases. Neurology 2004, 62, 411–413. [Google Scholar] [CrossRef] [PubMed]
  277. Okun, M.S.; McDonald, W.M.; DeLong, M.R. Refractory nonmotor symptoms in male patients with Parkinson disease due to testosterone deficiency: A common unrecognized comorbidity. Arch. Neurol. 2002, 59, 807–811. [Google Scholar] [CrossRef] [PubMed]
  278. Markianos, M.; Panas, M.; Kalfakis, N.; Vassilopoulos, D. Plasma testosterone in male patients with Huntington’s disease: Relations to severity of illness and dementia. Ann. Neurol. 2005, 57, 520–525. [Google Scholar] [CrossRef] [PubMed]
  279. Bove, R.; Musallam, A.; Healy, B.C.; Raghavan, K.; Glanz, B.I.; Bakshi, R.; Weiner, H.; De Jager, P.L.; Miller, K.K.; Chitnis, T. Low testosterone is associated with disability in men with multiple sclerosis. Mult. Scler. 2014, 20, 1584–1592. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  280. Trenova, A.G.; Slavov, G.S.; Manova, M.G.; Kostadinova, II; Vasileva, T.V. Female sex hormones and cytokine secretion in women with multiple sclerosis. Neurol. Res. 2013, 35, 95–99. [Google Scholar] [CrossRef] [PubMed]
  281. Zakrzewska-Pniewska, B.; Golebiowski, M.; Zajda, M.; Szeszkowski, W.; Podlecka-Pietowska, A.; Nojszewska, M. Sex hormone patterns in women with multiple sclerosis as related to disease activity—A pilot study. Neurol. I Neurochir. Polska 2011, 45, 536–542. [Google Scholar] [CrossRef]
  282. Baker, L.D.; Sambamurti, K.; Craft, S.; Cherrier, M.; Raskind, M.A.; Stanczyk, F.Z.; Plymate, S.R.; Asthana, S. 17beta-estradiol reduces plasma Abeta40 for HRT-naive postmenopausal women with Alzheimer disease: A preliminary study. Am. J. Geriatr. Psychiatry 2003, 11, 239–244. [Google Scholar] [CrossRef] [PubMed]
  283. Valen-Sendstad, A.; Engedal, K.; Stray-Pedersen, B.; Group, A.S.; Strobel, C.; Barnett, L.; Meyer, N.; Nurminemi, M. Effects of hormone therapy on depressive symptoms and cognitive functions in women with Alzheimer disease: A 12 month randomized, double-blind, placebo-controlled study of low-dose estradiol and norethisterone. Am. J. Geriatr. Psychiatry 2010, 18, 11–20. [Google Scholar] [CrossRef] [PubMed]
  284. Cherrier, M.M.; Matsumoto, A.M.; Amory, J.K.; Asthana, S.; Bremner, W.; Peskind, E.R.; Raskind, M.A.; Craft, S. Testosterone improves spatial memory in men with Alzheimer disease and mild cognitive impairment. Neurology 2005, 64, 2063–2068. [Google Scholar] [CrossRef] [PubMed]
  285. Lu, P.H.; Masterman, D.A.; Mulnard, R.; Cotman, C.; Miller, B.; Yaffe, K.; Reback, E.; Porter, V.; Swerdloff, R.; Cummings, J.L. Effects of testosterone on cognition and mood in male patients with mild Alzheimer disease and healthy elderly men. Arch. Neurol. 2006, 63, 177–185. [Google Scholar] [CrossRef] [PubMed]
  286. Moffat, S.D.; Zonderman, A.B.; Metter, E.J.; Kawas, C.; Blackman, M.R.; Harman, S.M.; Resnick, S.M. Free testosterone and risk for Alzheimer disease in older men. Neurology 2004, 62, 188–193. [Google Scholar] [CrossRef] [PubMed]
  287. Martins, R.; Carruthers, M. Testosterone as the missing link between pesticides, Alzheimer disease, and Parkinson disease. JAMA Neurol. 2014, 71, 1189–1190. [Google Scholar] [CrossRef] [PubMed]
  288. Okun, M.S.; Fernandez, H.H.; Rodriguez, R.L.; Romrell, J.; Suelter, M.; Munson, S.; Louis, E.D.; Mulligan, T.; Foster, P.S.; Shenal, B.V.; et al. Testosterone therapy in men with Parkinson disease: Results of the test-PD study. Arch. Neurol. 2006, 63, 729–735. [Google Scholar] [CrossRef] [PubMed]
  289. Okun, M.S.; Walter, B.L.; McDonald, W.M.; Tenover, J.L.; Green, J.; Juncos, J.L.; DeLong, M.R. Beneficial effects of testosterone replacement for the nonmotor symptoms of Parkinson disease. Arch. Neurol. 2002, 59, 1750–1753. [Google Scholar] [CrossRef] [PubMed]
  290. Gold, S.M.; Voskuhl, R.R. Estrogen and testosterone therapies in multiple sclerosis. Prog. Brain Res. 2009, 175, 239–251. [Google Scholar] [PubMed] [Green Version]
  291. Kurth, F.; Luders, E.; Sicotte, N.L.; Gaser, C.; Giesser, B.S.; Swerdloff, R.S.; Montag, M.J.; Voskuhl, R.R.; Mackenzie-Graham, A. Neuroprotective effects of testosterone treatment in men with multiple sclerosis. Neuroimage Clin. 2014, 4, 454–460. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  292. Sicotte, N.L.; Giesser, B.S.; Tandon, V.; Klutch, R.; Steiner, B.; Drain, A.E.; Shattuck, D.W.; Hull, L.; Wang, H.J.; Elashoff, R.M.; et al. Testosterone treatment in multiple sclerosis: A pilot study. Arch. Neurol. 2007, 64, 683–688. [Google Scholar] [CrossRef] [PubMed]
  293. Vukusic, S.; Ionescu, I.; El-Etr, M.; Schumacher, M.; Baulieu, E.E.; Cornu, C.; Confavreux, C.; Prevention of Post-Partum Relapses with Progestin; Estradiol in Multiple Sclerosis Study Group. The prevention of post-partum relapses with progestin and estradiol in multiple sclerosis (POPART’MUS) trial: Rationale, objectives and state of advancement. J. Neurol. Sci. 2009, 286, 114–118. [Google Scholar] [CrossRef] [PubMed]
  294. Palaszynski, K.M.; Loo, K.K.; Ashouri, J.F.; Liu, H.B.; Voskuhl, R.R. Androgens are protective in experimental autoimmune encephalomyelitis: Implications for multiple sclerosis. J. Neuroimmunol. 2004, 146, 144–152. [Google Scholar] [CrossRef] [PubMed]
  295. Page, S.T.; Plymate, S.R.; Bremner, W.J.; Matsumoto, A.M.; Hess, D.L.; Lin, D.W.; Amory, J.K.; Nelson, P.S.; Wu, J.D. Effect of medical castration on CD4+ CD25+ T cells, CD8+ T cell IFN-gamma expression, and NK cells: A physiological role for testosterone and/or its metabolites. Am. J. Physiology. Endocrinol. Metab. 2006, 290, E856–E863. [Google Scholar] [CrossRef] [PubMed]
  296. Malkin, C.J.; Pugh, P.J.; Jones, R.D.; Kapoor, D.; Channer, K.S.; Jones, T.H. The effect of testosterone replacement on endogenous inflammatory cytokines and lipid profiles in hypogonadal men. J. Clin. Endocrinol. Metab. 2004, 89, 3313–3318. [Google Scholar] [CrossRef] [PubMed]
  297. Liva, S.M.; Voskuhl, R.R. Testosterone acts directly on CD4+ T lymphocytes to increase IL-10 production. J. Immunol. (Baltimore, Md. 1950) 2001, 167, 2060–2067. [Google Scholar] [CrossRef]
  298. Simpkins, J.W.; Singh, M.; Brock, C.; Etgen, A.M. Neuroprotection and estrogen receptors. Neuroendocrinology 2012, 96, 119–130. [Google Scholar] [CrossRef] [PubMed]
  299. Bryant, D.N.; Dorsa, D.M. Roles of estrogen receptors alpha and beta in sexually dimorphic neuroprotection against glutamate toxicity. Neuroscience 2010, 170, 1261–1269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  300. Deecher, D.C.; Daoud, P.; Bhat, R.A.; O’Connor, L.T. Endogenously expressed estrogen receptors mediate neuroprotection in hippocampal cells (HT22). J. Cell. Biochem. 2005, 95, 302–312. [Google Scholar] [CrossRef] [PubMed]
  301. Labombarda, F.; Ghoumari, A.M.; Liere, P.; De Nicola, A.F.; Schumacher, M.; Guennoun, R. Neuroprotection by steroids after neurotrauma in organotypic spinal cord cultures: A key role for progesterone receptors and steroidal modulators of Gaba(A) receptors. Neuropharmacology 2013, 71, 46–55. [Google Scholar] [CrossRef] [PubMed]
  302. Liu, A.; Margaill, I.; Zhang, S.; Labombarda, F.; Coqueran, B.; Delespierre, B.; Liere, P.; Marchand-Leroux, C.; O’Malley, B.W.; Lydon, J.P.; et al. Progesterone receptors: A key for neuroprotection in experimental stroke. Endocrinology 2012, 153, 3747–3757. [Google Scholar] [CrossRef] [PubMed]
  303. Dubal, D.B.; Wise, P.M. Neuroprotective effects of estradiol in middle-aged female rats. Endocrinology 2001, 142, 43–48. [Google Scholar] [CrossRef] [PubMed]
  304. Nuzzo, M.T.; Marino, M. Estrogen/huntingtin: A novel pathway involved in neuroprotection. Neural Regen. Res. 2016, 11, 402–403. [Google Scholar] [PubMed]
  305. Shen, D.; Tian, X.; Zhang, B.; Song, R. Mechanistic evaluation of neuroprotective effect of estradiol on rotenone and 6-ohda induced Parkinson’s disease. Pharmacol. Rep. PR 2017, 69, 1178–1185. [Google Scholar] [CrossRef] [PubMed]
  306. Smith, K.; Dahodwala, N. Neuroprotection by sex steroid hormones in Parkinson’s disease (P3.068). Neurology 2014, 82. [Google Scholar]
  307. Zárate, S.; Stevnsner, T.; Gredilla, R. Role of estrogen and other sex hormones in brain aging. Neuroprotection and DNA repair. Front. Aging Neurosci. 2017, 9, 430. [Google Scholar] [CrossRef] [PubMed]
  308. Barker, J.M.; Galea, L.A. Repeated estradiol administration alters different aspects of neurogenesis and cell death in the hippocampus of female, but not male, rats. Neuroscience 2008, 152, 888–902. [Google Scholar] [CrossRef] [PubMed]
  309. Spencer, J.L.; Waters, E.M.; Romeo, R.D.; Wood, G.E.; Milner, T.A.; McEwen, B.S. Uncovering the mechanisms of estrogen effects on hippocampal function. Front. Neuroendocr. 2008, 29, 219–237. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  310. Knoll, J.G.; Wolfe, C.A.; Tobet, S.A. Estrogen modulates neuronal movements within the developing preoptic area/anterior hypothalamus. Eur. J. Neurosci. 2007, 26, 1091–1099. [Google Scholar] [CrossRef] [PubMed]
  311. Tiwari-Woodruff, S.; Morales, L.B.J.; Lee, R.; Voskuhl, R.R. Differential neuroprotective and antiinflammatory effects of estrogen receptor (ER)α and ERβ ligand treatment. Proc. Natl. Acad. Sci. USA 2007, 104, 14813–14818. [Google Scholar] [CrossRef] [PubMed]
  312. Zhang, Q.-G.; Wang, R.; Tang, H.; Dong, Y.; Chan, A.; Sareddy, G.R.; Vadlamudi, R.K.; Brann, D.W. Brain-derived estrogen exerts anti-inflammatory and neuroprotective actions in the rat hippocampus. Mol. Cell. Endocrinol. 2014, 389, 84–91. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  313. Shivers, K.-Y.; Amador, N.; Abrams, L.; Hunter, D.; Jenab, S.; Quiñones-Jenab, V. Estrogen alters baseline and inflammatory-induced cytokine levels independent from hypothalamic–pituitary–adrenal axis activity. Cytokine 2015, 72, 121–129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  314. Brown, C.M.; Mulcahey, T.A.; Filipek, N.C.; Wise, P.M. Production of proinflammatory cytokines and chemokines during neuroinflammation: Novel roles for estrogen receptors α and β. Endocrinology 2010, 151, 4916–4925. [Google Scholar] [CrossRef] [PubMed]
  315. Ahn, J.-Y. Neuroprotection signaling of nuclear Akt in neuronal cells. Exp. Neurobiol. 2014, 23, 200–206. [Google Scholar] [CrossRef] [PubMed]
  316. Wang, L.; Zhou, K.; Fu, Z.; Yu, D.; Huang, H.; Zang, X.; Mo, X. Brain development and Akt signaling: The crossroads of signaling pathway and neurodevelopmental diseases. J. Mol. Neurosci. 2017, 61, 379–384. [Google Scholar] [CrossRef] [PubMed]
  317. Itoh, Y.; Higuchi, M.; Oishi, K.; Kishi, Y.; Okazaki, T.; Sakai, H.; Miyata, T.; Nakajima, K.; Gotoh, Y. PDK1–Akt pathway regulates radial neuronal migration and microtubules in the developing mouse neocortex. Proc. Natl. Acad. Sci. USA 2016, 113, E2955–E2964. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  318. Liot, G.; Gabriel, C.; Cacquevel, M.; Ali, C.; MacKenzie, E.T.; Buisson, A.; Vivien, D. Neurotrophin-3-induced PI-3 kinase/Akt signaling rescues cortical neurons from apoptosis. Exp. Neurol. 2004, 187, 38–46. [Google Scholar] [CrossRef] [PubMed]
  319. Gross, C.; Bassell, G.J. Neuron-specific regulation of class I PI3K catalytic subunits and their dysfunction in brain disorders. Front. Mol. Neurosci. 2014, 7, 12. [Google Scholar] [CrossRef] [PubMed]
  320. Guo, W.; Jiang, H.; Gray, V.; Dedhar, S.; Rao, Y. Role of the integrin-linked kinase (ILK) in determining neuronal polarity. Dev. Boil. 2007, 306, 457–468. [Google Scholar] [CrossRef] [PubMed]
  321. Hussain, R.; Macklin, W.B. Integrin-linked kinase (ILK) deletion disrupts oligodendrocyte development by altering cell cycle. J. Neurosci. 2017, 37, 397–412. [Google Scholar] [CrossRef] [PubMed]
  322. Pereira, J.A.; Benninger, Y.; Baumann, R.; Gonçalves, A.F.; Özçelik, M.; Thurnherr, T.; Tricaud, N.; Meijer, D.; Fässler, R.; Suter, U.; et al. Integrin-linked kinase is required for radial sorting of axons and schwann cell remyelination in the peripheral nervous system. J. Cell Boil. 2009, 185, 147–161. [Google Scholar] [CrossRef] [PubMed]
  323. Soltani, M.H.; Pichardo, R.; Song, Z.; Sangha, N.; Camacho, F.; Satyamoorthy, K.; Sangueza, O.P.; Setaluri, V. Microtubule-associated protein 2, a marker of neuronal differentiation, induces mitotic defects, inhibits growth of melanoma cells, and predicts metastatic potential of cutaneous melanoma. Am. J. Pathol. 2005, 166, 1841–1850. [Google Scholar] [CrossRef]
  324. Johnson, G.V.; Jope, R.S. The role of microtubule-associated protein 2 (map-2) in neuronal growth, plasticity, and degeneration. J. Neurosci. Res. 1992, 33, 505–512. [Google Scholar] [CrossRef] [PubMed]
  325. Chen, W.; Lu, N.; Ding, Y.; Wang, Y.; Chan, L.T.; Wang, X.; Gao, X.; Jiang, S.; Liu, K. Rapamycin-resistant mTOR activity is required for sensory axon regeneration induced by a conditioning lesion. eNeuro 2016, 3. [Google Scholar] [CrossRef] [PubMed]
  326. Maiese, K. Driving neural regeneration through the mammalian target of rapamycin. Neural Regen. Res. 2014, 9, 1413. [Google Scholar] [CrossRef] [PubMed]
  327. Takei, N.; Nawa, H. mTOR signaling and its roles in normal and abnormal brain development. Front. Mol. Neurosci. 2014, 7, 28. [Google Scholar] [CrossRef] [PubMed]
  328. Blázquez, E.; Velázquez, E.; Hurtado-Carneiro, V.; Ruiz-Albusac, J.M. Insulin in the brain: Its pathophysiological implications for states related with central insulin resistance, type 2 diabetes and Alzheimer’s disease. Front. Endocrinol. 2014, 5, 161. [Google Scholar] [CrossRef] [PubMed]
  329. Hoyer, S. Senile dementia and Alzheimer’s disease. Brain blood flow and metabolism. Prog. Neuro-Psychopharmacol. Boil. Psychiatry 1986, 10, 447–478. [Google Scholar] [CrossRef]
  330. Misiak, B.; Leszek, J.; Kiejna, A. Metabolic syndrome, mild cognitive impairment and Alzheimer’s disease—The emerging role of systemic low-grade inflammation and adiposity. Brain Res. Bull. 2012, 89, 144–149. [Google Scholar] [CrossRef] [PubMed]
  331. Gaspar, J.M.; Baptista, F.I.; Macedo, M.P.; Ambrósio, A.N.F. Inside the diabetic brain: Role of different players involved in cognitive decline. ACS Chem. Neurosci. 2015, 7, 131–142. [Google Scholar] [CrossRef] [PubMed]
  332. de la Monte, S.M.; Wands, J.R. Alzheimer’s disease is type 3 diabetes—Evidence reviewed. J. Diabetes Sci. Technol. 2008, 2, 1101–1113. [Google Scholar] [CrossRef] [PubMed]
  333. de la Monte, S.M.; Tong, M.; Bowling, N.; Moskal, P. si-RNA inhibition of brain insulin or insulin-like growth factor receptors causes developmental cerebellar abnormalities: Relevance to fetal alcohol spectrum disorder. Mol. Brain 2011, 4, 13. [Google Scholar] [CrossRef] [PubMed]
  334. Lee, S.; Tong, M.; Hang, S.; Deochand, C.; de la Monte, S. CSF and brain indices of insulin resistance, oxidative stress and neuro-inflammation in early versus late Alzheimer’s disease. J. Alzheimer’s Dis. Park. 2013, 3, 128. [Google Scholar]
  335. Shuvaev, V.V.; Laffont, I.; Serot, J.-M.; Fujii, J.; Taniguchi, N.; Siest, G. Increased protein glycation in cerebrospinal fluid of Alzheimer’s disease2. Neurobiol. Aging 2001, 22, 397–402. [Google Scholar] [CrossRef]
  336. Benedict, C.; Frey II, W.H.; Schiöth, H.B.; Schultes, B.; Born, J.; Hallschmid, M. Intranasal insulin as a therapeutic option in the treatment of cognitive impairments. Exp. Gerontol. 2011, 46, 112–115. [Google Scholar] [CrossRef] [PubMed]
  337. Reger, M.; Watson, G.; Frey Ii, W.; Baker, L.; Cholerton, B.; Keeling, M.; Belongia, D.; Fishel, M.; Plymate, S.; Schellenberg, G. Effects of intranasal insulin on cognition in memory-impaired older adults: Modulation by APOE genotype. Neurobiol. Aging 2006, 27, 451–458. [Google Scholar] [CrossRef] [PubMed]
  338. White, M.F. Insulin signaling in health and disease. Science (New York) 2003, 302, 1710–1711. [Google Scholar] [CrossRef] [PubMed]
  339. Boura-Halfon, S.; Zick, Y. Phosphorylation of IRS proteins, insulin action, and insulin resistance. Am. J. Physiol. Endocrinol. Metab. 2009, 296, E581–E591. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  340. Giordano, G.; Klintworth, H.; Kavanagh, T.; Costa, L. Apoptosis induced by domoic acid in mouse cerebellar granule neurons involves activation of p38 and JNK MAP kinases. Neurochem. Int. 2008, 52, 1100–1105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  341. Andreozzi, F.; Laratta, E.; Procopio, C.; Hribal, M.L.; Sciacqua, A.; Perticone, M.; Miele, C.; Perticone, F.; Sesti, G. Interleukin-6 impairs the insulin signaling pathway, promoting production of nitric oxide in human umbilical vein endothelial cells. Mol. Cell. Boil. 2007, 27, 2372–2383. [Google Scholar] [CrossRef] [PubMed]
  342. Bomfim, T.R.; Forny-Germano, L.; Sathler, L.B.; Brito-Moreira, J.; Houzel, J.-C.; Decker, H.; Silverman, M.A.; Kazi, H.; Melo, H.M.; McClean, P.L. An anti-diabetes agent protects the mouse brain from defective insulin signaling caused by Alzheimer’s disease–associated Aβ oligomers. J. Clin. Investig. 2012, 122, 1339–1353. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  343. Dineley, K.T.; Westerman, M.; Bui, D.; Bell, K.; Ashe, K.H.; Sweatt, J.D. Β-amyloid activates the mitogen-activated protein kinase cascade via hippocampal α7 nicotinic acetylcholine receptors: In vitro and in vivo mechanisms related to Alzheimer’s disease. J. Neurosci. 2001, 21, 4125–4133. [Google Scholar] [CrossRef] [PubMed]
  344. Talbot, K.; Wang, H.-Y.; Kazi, H.; Han, L.-Y.; Bakshi, K.P.; Stucky, A.; Fuino, R.L.; Kawaguchi, K.R.; Samoyedny, A.J.; Wilson, R.S. Demonstrated brain insulin resistance in Alzheimer’s disease patients is associated with IGF-1 resistance, IRS-1 dysregulation, and cognitive decline. J. Clin. Investig. 2012, 122, 1316–1338. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  345. Unger, J.W.; Livingston, J.N.; Moss, A.M. Insulin receptors in the central nervous system: Localization, signalling mechanisms and functional aspects. Prog. Neurobiol. 1991, 36, 343–362. [Google Scholar] [CrossRef]
  346. Moroo, I.; Yamada, T.; Makino, H.; Tooyama, I.; McGeer, P.; McGeer, E.; Hirayama, K. Loss of insulin receptor immunoreactivity from the substantia nigra pars compacta neurons in Parkinson’s disease. Acta Neuropathol. 1994, 87, 343–348. [Google Scholar] [CrossRef] [PubMed]
  347. Takahashi, M.; Yamada, T.; Tooyama, I.; Moroo, I.; Kimura, H.; Yamamoto, T.; Okada, H. Insulin receptor mRNA in the substantia nigra in Parkinson’s disease. Neurosci. Lett. 1996, 204, 201–204. [Google Scholar] [CrossRef]
  348. Bowen, B.C.; Block, R.E.; Sanchez-Ramos, J.; Pattany, P.M.; Lampman, D.A.; Murdoch, J.B.; Quencer, R.M. Proton MR spectroscopy of the brain in 14 patients with Parkinson disease. Am. J. Neuroradiol. 1995, 16, 61–68. [Google Scholar] [PubMed]
  349. Hu, M.; Taylor-Robinson, S.D.; Chaudhuri, K.R.; Bell, J.D.; Labbe, C.; Cunningham, V.; Koepp, M.; Hammers, A.; Morris, R.; Turjanski, N. Cortical dysfunction in non-demented Parkinson’s disease patients: A combined 31P-MRS and 18FDG-PET study. Brain J. Neurol. 2000, 123, 340–352. [Google Scholar] [CrossRef]
  350. Lalić, N.M.; Marić, J.; Svetel, M.; Jotić, A.; Stefanova, E.; Lalić, K.; Dragašević, N.; Miličić, T.; Lukić, L.; Kostić, V.S. Glucose homeostasis in Huntington disease: Abnormalities in insulin sensitivity and early-phase insulin secretion. Arch. Neurol. 2008, 65, 476–480. [Google Scholar] [CrossRef] [PubMed]
  351. Montojo, M.T.; Aganzo, M.; González, N. Huntington’s disease and diabetes: Chronological sequence of its association. J. Huntington’s Dis. 2017, 6, 179–188. [Google Scholar] [CrossRef] [PubMed]
  352. Viollet, B.; Guigas, B.; Garcia, N.S.; Leclerc, J.; Foretz, M.; Andreelli, F. Cellular and molecular mechanisms of metformin: An overview. Clin. Sci. 2012, 122, 253–270. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  353. Patrone, C.; Eriksson, O.; Lindholm, D. Diabetes drugs and neurological disorders: New views and therapeutic possibilities. Lancet Diabetes Endocrinol. 2014, 2, 256–262. [Google Scholar] [CrossRef]
  354. Bayliss, J.A.; Lemus, M.B.; Santos, V.V.; Deo, M.; Davies, J.S.; Kemp, B.E.; Elsworth, J.D.; Andrews, Z.B. Metformin prevents nigrostriatal dopamine degeneration independent of AMPK activation in dopamine neurons. PLoS ONE 2016, 11, e0159381. [Google Scholar] [CrossRef] [PubMed]
  355. Ismaiel, A.A.; Espinosa-Oliva, A.M.; Santiago, M.; García-Quintanilla, A.; Oliva-Martín, M.J.; Herrera, A.J.; Venero, J.L.; de Pablos, R.M. Metformin, besides exhibiting strong in vivo anti-inflammatory properties, increases MPTP-induced damage to the nigrostriatal dopaminergic system. Toxicol. Appl. Pharmacol. 2016, 298, 19–30. [Google Scholar] [CrossRef] [PubMed]
  356. Kang, H.; Khang, R.; Ham, S.; Jeong, G.R.; Kim, H.; Jo, M.; Lee, B.D.; Lee, Y.I.; Jo, A.; Park, C. Activation of the ATF2/CREB-PGC-1α pathway by metformin leads to dopaminergic neuroprotection. Oncotarget 2017, 8, 48603. [Google Scholar] [CrossRef] [PubMed]
  357. Chen, F.; Dong, R.R.; Zhong, K.L.; Ghosh, A.; Tang, S.S.; Long, Y.; Hu, M.; Miao, M.X.; Liao, J.M.; Sun, H.B. Antidiabetic drugs restore abnormal transport of amyloid-β across the blood–brain barrier and memory impairment in db/db mice. Neuropharmacology 2016, 101, 123–136. [Google Scholar] [CrossRef] [PubMed]
  358. Wahlqvist, M.L.; Lee, M.-S.; Hsu, C.-C.; Chuang, S.-Y.; Lee, J.-T.; Tsai, H.-N. Metformin-inclusive sulfonylurea therapy reduces the risk of Parkinson’s disease occurring with type 2 diabetes in a Taiwanese population cohort. Park. Relat. Disord. 2012, 18, 753–758. [Google Scholar] [CrossRef] [PubMed]
  359. Ye, F.; Luo, Y.-J.; Xiao, J.; Yu, N.-W.; Yi, G. Impact of insulin sensitizers on the incidence of dementia: A meta-analysis. Dement. Geriatr. Cogn. Disord. 2016, 41, 251–260. [Google Scholar] [CrossRef] [PubMed]
  360. Koenig, A.M.; Mechanic-Hamilton, D.; Xie, S.X.; Combs, M.F.; Cappola, A.R.; Xie, L.; Detre, J.A.; Wolk, D.A.; Arnold, S.E. Effects of the insulin sensitizer metformin in Alzheimer’s disease: Pilot data from a randomized placebo-controlled crossover study. Alzheimer Dis. Assoc. Disord. 2017, 31, 107. [Google Scholar] [CrossRef] [PubMed]
  361. Hervás, D.; Fornés-Ferrer, V.; Gómez-Escribano, A.P.; Sequedo, M.D.; Peiró, C.; Millán, J.M.; Vázquez-Manrique, R.P. Metformin intake associates with better cognitive function in patients with Huntington’s disease. PLoS ONE 2017, 12, e0179283. [Google Scholar] [CrossRef] [PubMed]
  362. Hampel, H.; Mesulam, M.-M.; Cuello, A.C.; Farlow, M.R.; Giacobini, E.; Grossberg, G.T.; Khachaturian, A.S.; Vergallo, A.; Cavedo, E.; Snyder, P.J. The cholinergic system in the pathophysiology and treatment of Alzheimer’s disease. Brain J. Neurol. 2018, 141, 1917–1933. [Google Scholar] [CrossRef] [PubMed]
  363. Mesulam, M.M. Cholinergic circuitry of the human nucleus basalis and its fate in Alzheimer’s disease. J. Comp. Neurol. 2013, 521, 4124–4144. [Google Scholar] [CrossRef] [PubMed]
  364. Bowen, D.M.; Smith, C.B.; White, P.; Davison, A.N. Neurotransmitter-related enzymes and indices of hypoxia in senile dementia and other abiotrophies. Brain J. Neurol. 1976, 99, 459–496. [Google Scholar] [CrossRef]
  365. Davies, P.; Maloney, A. Selective loss of central cholinergic neurons in Alzheimer’s disease. Lancet 1976, 308, 1403. [Google Scholar] [CrossRef]
  366. Mesulam, M. A horseradish peroxidase method for the identification of the efferents of acetyl cholinesterase-containing neurons. J. Histochem. Cytochem. 1976, 24, 1281–1285. [Google Scholar] [CrossRef] [PubMed]
  367. Whitehouse, P.J.; Price, D.L.; Clark, A.W.; Coyle, J.T.; DeLong, M.R. Alzheimer disease: Evidence for selective loss of cholinergic neurons in the nucleus basalis. Ann. Neurol. Off. J. Am. Neurol. Assoc. Child Neurol. Soc. 1981, 10, 122–126. [Google Scholar] [CrossRef] [PubMed]
  368. Drachman, D.A.; Leavitt, J. Human memory and the cholinergic system: A relationship to aging? Arch. Neurol. 1974, 30, 113–121. [Google Scholar] [CrossRef] [PubMed]
  369. Summers, W.K.; Majovski, L.V.; Marsh, G.M.; Tachiki, K.; Kling, A. Oral tetrahydroaminoacridine in long-term treatment of senile dementia, Alzheimer type. N. Engl. J. Med. 1986, 315, 1241–1245. [Google Scholar] [CrossRef] [PubMed]
  370. Petersen, R.C. Scopolamine induced learning failures in man. Psychopharmacology 1977, 52, 283–289. [Google Scholar] [CrossRef] [PubMed]
  371. Izquierdo, I. Mechanism of action of scopolamine as an amnestic. Trends Pharmacol. Sci. 1989, 10, 175–177. [Google Scholar] [CrossRef]
  372. Lovestone, S.; Howard, R. Alzheimer’s disease: A treatment in sight? J. Neurol. Neurosurgery, Psychiatry 1995, 59, 566. [Google Scholar] [CrossRef]
  373. Massoud, F.; Gauthier, S. Update on the pharmacological treatment of Alzheimer’s disease. Curr. Neuropharmacol. 2010, 8, 69–80. [Google Scholar] [CrossRef] [PubMed]
  374. Hansen, R.A.; Gartlehner, G.; Webb, A.P.; Morgan, L.C.; Moore, C.G.; Jonas, D.E. Efficacy and safety of donepezil, galantamine, and rivastigmine for the treatment of Alzheimer’s disease: A systematic review and meta-analysis. Clin. Interv. Aging 2008, 3, 211. [Google Scholar] [PubMed]
  375. Richter, N.; Beckers, N.; Onur, O.A.; Dietlein, M.; Tittgemeyer, M.; Kracht, L.; Neumaier, B.; Fink, G.R.; Kukolja, J. Effect of cholinergic treatment depends on cholinergic integrity in early Alzheimer’s disease. Brain J. Neurol. 2018, 141, 903–915. [Google Scholar] [CrossRef] [PubMed]
  376. Schmitz, T.W.; Spreng, R.N.; Weiner, M.W.; Aisen, P.; Petersen, R.; Jack, C.R.; Jagust, W.; Trojanowki, J.Q.; Toga, A.W.; Beckett, L. Basal forebrain degeneration precedes and predicts the cortical spread of Alzheimer’s pathology. Nat. Commun. 2016, 7, 13249. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  377. Sigurdsson, E.M.; Permanne, B.; Soto, C.; Wisniewski, T.; Frangione, B. In vivo reversal of amyloid-β lesions in rat brain. J. Neuropathol. Exp. Neurol. 2000, 59, 11–17. [Google Scholar] [CrossRef] [PubMed]
  378. Morgan, D.; Diamond, D.M.; Gottschall, P.E.; Ugen, K.E.; Dickey, C.; Hardy, J.; Duff, K.; Jantzen, P.; DiCarlo, G.; Wilcock, D. Aβ peptide vaccination prevents memory loss in an animal model of Alzheimer’s disease. Nature 2000, 408, 982–985. [Google Scholar] [CrossRef] [PubMed]
  379. Janus, C.; Pearson, J.; McLaurin, J.; Mathews, P.M.; Jiang, Y.; Schmidt, S.D.; Chishti, M.A.; Horne, P.; Heslin, D.; French, J. Aβ peptide immunization reduces behavioural impairment and plaques in a model of Alzheimer’s disease. Nature 2000, 408, 979–982. [Google Scholar] [CrossRef] [PubMed]
  380. Schenk, D.; Barbour, R.; Dunn, W.; Gordon, G.; Grajeda, H.; Guido, T.; Hu, K.; Huang, J.; Johnson-Wood, K.; Khan, K. Immunization with amyloid-β attenuates Alzheimer-disease-like pathology in the PDAPP mouse. Nature 1999, 400, 173–177. [Google Scholar] [CrossRef] [PubMed]
  381. Vandenberghe, R.; Rinne, J.O.; Boada, M.; Katayama, S.; Scheltens, P.; Vellas, B.; Tuchman, M.; Gass, A.; Fiebach, J.B.; Hill, D. Bapineuzumab for mild to moderate Alzheimer’s disease in two global, randomized, phase 3 trials. Alzheimer’s Res. Ther. 2016, 8, 18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  382. Lacey, L.; Bobula, J.; Rüdell, K.; Alvir, J.; Leibman, C. Quality of life and utility measurement in a large clinical trial sample of patients with mild to moderate Alzheimer’s disease: Determinants and level of changes observed. Value Health 2015, 18, 638–645. [Google Scholar] [CrossRef] [PubMed]
  383. Honig, L.S.; Vellas, B.; Woodward, M.; Boada, M.; Bullock, R.; Borrie, M.; Hager, K.; Andreasen, N.; Scarpini, E.; Liu-Seifert, H. Trial of solanezumab for mild dementia due to Alzheimer’s disease. N. Engl. J. Med. 2018, 378, 321–330. [Google Scholar] [CrossRef] [PubMed]
  384. Siemers, E.; Sundell, K.; Carlson, C.; Case, M.; Sethuraman, G.; Liu-Seifert, H.; Dowsett, S.; Pontecorvo, M.; Dean, R.; Demattos, R. Phase 3 solanezumab trials: Secondary outcomes in mild Alzheimer’s disease patients. Alzheimer’s Dement. 1983, 12, 1–11. [Google Scholar] [CrossRef] [PubMed]
  385. Bateman, R.J.; Benzinger, T.L.; Berry, S.; Clifford, D.B.; Duggan, C.; Fagan, A.M.; Fanning, K.; Farlow, M.R.; Hassenstab, J.; McDade, E.M. The DIAN-TU next generation Alzheimer’s prevention trial: Adaptive design and disease progression model. Alzheimer’s Dement. J. Alzheimer’s Assoc. 2017, 13, 8–19. [Google Scholar] [CrossRef] [PubMed]
  386. DeMattos, R.B.; Lu, J.; Tang, Y.; Racke, M.M.; DeLong, C.A.; Tzaferis, J.A.; Hole, J.T.; Forster, B.M.; McDonnell, P.C.; Liu, F. A plaque-specific antibody clears existing β-amyloid plaques in Alzheimer’s disease mice. Neuron 2012, 76, 908–920. [Google Scholar] [CrossRef] [PubMed]
  387. Irizarry, M.C.; Sims, J.R.; Lowe, S.L.; Nakano, M.; Hawdon, A.; Willis, B.A.; Gonzales, C.; Liu, P.; Fujimoto, S.; Dean, R.A. Safety, pharmacokinetics (PK), and florbetapir F-18 positron emission tomography (PET) after multiple dose administration of LY3002813, a β-amyloid plaque-specific antibody, in Alzheimer’s disease (AD). Alzheimer’s Dement. J. Alzheimer’s Assoc. 2016, 12, P352–P353. [Google Scholar] [CrossRef]
  388. Bohrmann, B.; Baumann, K.; Benz, J.; Gerber, F.; Huber, W.; Knoflach, F.; Messer, J.; Oroszlan, K.; Rauchenberger, R.; Richter, W.F. Gantenerumab: A novel human anti-Aβ antibody demonstrates sustained cerebral amyloid-β binding and elicits cell-mediated removal of human amyloid-β. J. Alzheimer’s Dis. 2012, 28, 49–69. [Google Scholar] [CrossRef] [PubMed]
  389. Blaettler, T.; Smith, J.; Smith, J.; Paul, R.; Asnaghi, V.; Fuji, R.; Quartino, A.; Honigberg, L.; Rabbia, M.A.; Yule, S. Clinical trial design of cread: A randomized, double-blind, placebo-controlled, parallel-group phase 3 study to evaluate crenezumab treatment in patients with prodromal-to-mild Alzheimer’s disease. Alzheimer’s Dement. J. Alzheimer’s Assoc. 2016, 12, P609. [Google Scholar] [CrossRef]
  390. Adolfsson, O.; Pihlgren, M.; Toni, N.; Varisco, Y.; Buccarello, A.L.; Antoniello, K.; Lohmann, S.; Piorkowska, K.; Gafner, V.; Atwal, J.K. An effector-reduced anti-β-amyloid (Aβ) antibody with unique Aβ binding properties promotes neuroprotection and glial engulfment of Aβ. J. Neurosci. 2012, 32, 9677–9689. [Google Scholar] [CrossRef] [PubMed]
  391. Tucker, S.; Möller, C.; Tegerstedt, K.; Lord, A.; Laudon, H.; Sjödahl, J.; Söderberg, L.; Spens, E.; Sahlin, C.; Waara, E.R. The murine version of BAN2401 (mAb158) selectively reduces amyloid-β protofibrils in brain and cerebrospinal fluid of tg-arcswe mice. J. Alzheimer’s Dis. 2015, 43, 575–588. [Google Scholar] [CrossRef] [PubMed]
  392. Logovinsky, V.; Satlin, A.; Lai, R.; Swanson, C.; Kaplow, J.; Osswald, G.; Basun, H.; Lannfelt, L. Safety and tolerability of BAN2401-a clinical study in Alzheimer’s disease with a protofibril selective Aβ antibody. Alzheimer’s Res. Ther. 2016, 8, 14. [Google Scholar] [CrossRef] [PubMed]
  393. Satlin, A.; Wang, J.; Logovinsky, V.; Berry, S.; Swanson, C.; Dhadda, S.; Berry, D.A. Design of a bayesian adaptive phase 2 proof-of-concept trial for BAN2401, a putative disease-modifying monoclonal antibody for the treatment of Alzheimer’s disease. Alzheimer’s Dement. Transl. Res. Clin. Interv. 2016, 2, 1–12. [Google Scholar] [CrossRef] [PubMed]
  394. Sevigny, J.; Chiao, P.; Bussière, T.; Weinreb, P.H.; Williams, L.; Maier, M.; Dunstan, R.; Salloway, S.; Chen, T.; Ling, Y. The antibody aducanumab reduces Aβ plaques in Alzheimer’s disease. Nature 2016, 537, 50–56. [Google Scholar] [CrossRef] [PubMed]
  395. Kastanenka, K.V.; Bussiere, T.; Shakerdge, N.; Qian, F.; Weinreb, P.H.; Rhodes, K.; Bacskai, B.J. Immunotherapy with aducanumab restores calcium homeostasis in Tg2576 mice. J. Neurosci. 2016, 36, 12549–12558. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Major Neurodegenerative diseases, their associated regions, and current therapeutic interventions. Left panel: Brain disorders are color and shown in representative areas of the brain. Right panel: current pharmacological treatments and their areas of activity within the brain. Abbreviations: Basal ganglion (BG), Brain Stem (BS), Cerebellum (Crbl), Corpus callosum (CC), Cortex (Cx), Hippocampus (Hp), Striatum (St), Substantia Nigra (SN).
Figure 1. Major Neurodegenerative diseases, their associated regions, and current therapeutic interventions. Left panel: Brain disorders are color and shown in representative areas of the brain. Right panel: current pharmacological treatments and their areas of activity within the brain. Abbreviations: Basal ganglion (BG), Brain Stem (BS), Cerebellum (Crbl), Corpus callosum (CC), Cortex (Cx), Hippocampus (Hp), Striatum (St), Substantia Nigra (SN).
Brainsci 08 00177 g001
Figure 2. The interplay of neurotrophic factors, steroids, and intra-cellular signaling, in growth and differentiation of neural tissue. Cells sense different cues in the extracellular environment through membranous receptors, and these changes are communicated downstream through cascades of protein/cytoplasmic factor activation and inactivation. Many of these changes go through Protein kinase B or AKT, and mammalian target of rapamycin (mTOR), which further stimulate ribosomal proteins, i.e., 4EBP and S6 kinases, as well as PKC and SG1K. Stimulation of ribosomal proteins leads to protein synthesis, proliferation, growth, and differentiation; whilst PKC and SGK1 activation suppresses apoptotic pathways and improves survival. Abbreviations: BDNS (brain-derived neurotrophic factor), NT3 (Netrins 3), TGF (transforming growth factor), NGF (nerve growth factor), EGF (epithelial growth factor), CTNF (ciliary neurotrophic factor), GPCR (G protein-coupled receptor), Pi3K (phosphoinositide 3 kinase), Akt (protein kinase b), mTORC1 (mammalian target of rapamycin complex1), mTORC2 (mammalian target of rapamycin complex 2), 4EBP (eukaryotic initiation factor 4 binding protein), S6K (ribosomal protein S6 kinase), PKC (protein kinase c), SGK1 (glucocorticoid regulated kinase 1).
Figure 2. The interplay of neurotrophic factors, steroids, and intra-cellular signaling, in growth and differentiation of neural tissue. Cells sense different cues in the extracellular environment through membranous receptors, and these changes are communicated downstream through cascades of protein/cytoplasmic factor activation and inactivation. Many of these changes go through Protein kinase B or AKT, and mammalian target of rapamycin (mTOR), which further stimulate ribosomal proteins, i.e., 4EBP and S6 kinases, as well as PKC and SG1K. Stimulation of ribosomal proteins leads to protein synthesis, proliferation, growth, and differentiation; whilst PKC and SGK1 activation suppresses apoptotic pathways and improves survival. Abbreviations: BDNS (brain-derived neurotrophic factor), NT3 (Netrins 3), TGF (transforming growth factor), NGF (nerve growth factor), EGF (epithelial growth factor), CTNF (ciliary neurotrophic factor), GPCR (G protein-coupled receptor), Pi3K (phosphoinositide 3 kinase), Akt (protein kinase b), mTORC1 (mammalian target of rapamycin complex1), mTORC2 (mammalian target of rapamycin complex 2), 4EBP (eukaryotic initiation factor 4 binding protein), S6K (ribosomal protein S6 kinase), PKC (protein kinase c), SGK1 (glucocorticoid regulated kinase 1).
Brainsci 08 00177 g002
Table 1. Summary of novel treatment strategies for neurodegenerative diseases.
Table 1. Summary of novel treatment strategies for neurodegenerative diseases.
StrategyModelDrugsDoseResultsReference
Inhibiting protein aggregatesRatiAβ5 (Chaperon)200 nmol
  • Reduced size and number of cerebral amyloid plaques in AD
[377]
Disaggregating misfolded proteinsMouseHsp104-
  • Degradation of mutant HTT
[171]
MouseGeldanamycin1 or 10 µg/kg
  • Inhibition of HSP90
  • Induction of HSP70
  • Neuroprotection in MPTP induced PD
[177]
ImmunomodulationMouseAmyloid-β-Peptide-
  • Reduction of Aβ deposits in the frontal cortex and hippocampus in HD model
  • Improved cognitive function
  • Reduced astrogliosis and disease progression
[378,379,380]
MouseResveratrol50 mg/kg
  • Reduced glial activation in PD model
  • Reduced ILs
  • Improved TH expression
[195]
MouseTanshinone IIA25 mg/kg
  • Reduced degeneration of nigrostriatal DA neurons
  • Increased striatal DA content
[197]
MouseTanshinone I10 mg/kg
  • Reduced expression of pro-inflammatory factors
  • Improved motor function and striatal neurotransmitters
[198]
Mouse1H7, 5C1, 5D12-
  • Reduced α-synuclein accumulation
  • Reduced synaptic and axonal pathology in PD model
  • Improved cognitive function
[204]
MousemAb47-
  • Reduced α-synuclein accumulation in PD model
  • Improved motor function
[205]
Induction of AutophagyDrosophilaRapamycin1 µM
  • Enhanced clearance of pyroglutamine and polyalanine proteins
  • Enhanced clearance of tau protein and decreased tau toxicity
[223]
VerapamilMice25 mg/kg
  • Suppressed hepatosteatosis
  • Reduced obesity-induced cytosolic calcium in liver
  • Restored autophagic flux
[225]
Rapamycin ester (CCI-779)Mice20 mg/kg
  • Enhanced clearance of mutant huntingtin
  • Improved behavioral tasks
[102]
MetforminMice2 mg/mL of drinking water
  • Prolonged survival time
  • Decreased hindlimb clasping time
[229]
NilotinibMice10 mg/kg
  • Reduced brain and peripheral α-synuclein and p-tau
  • Improved immune profile
[231]
NilotinibMice-
  • Increased β-amyloid clearance
  • Reduced inflammation
  • Improved immune profile
[232]
Table 2. Summary of monoclonal antibodies under clinical trials.
Table 2. Summary of monoclonal antibodies under clinical trials.
AntibodyIgG SubtypeSpecificityDoseReference
BapineuzumabIgG1 AAB-001 (humanized mouse 3D6)Aβ 1–5 (helical, N-terminal D sensitive)Phase I: 12-month 0.5, 1.5, or 5 mg/kg
Phase II: 18-month 0.15, 0.5, 1, or 2 mg/kg
Phase III: 18-month 0.5 mg/kg 1.0 mg/kg
[381,382]
SolanezumabIgG1 (humanized mouse [265])Aβ 16–26 accessible only on monomeric Aβ400 mg every week for 76 weeks Phase III A4 400–1600 mg every 4 weeks for 240 weeks[383,384,385]
LY3002813IgG1 (humanized mouse mE8-IgG2a)pE3-Aβ0.1 mg/kg to 10 mg/kg, infused monthly up to four times, and a single subcutaneous injection against placebo for safety[386,387]
Gantenerumabb (RG1450, RO4909832) IgG1 (full human)Aβ 2–5 (−9) + 23–25 binds with subnanomolar affinity to a conformational epitope on Aβ fibrils. It binds both N-terminal and central amino acids of AβPhase III 225 mg SC[388]
CrenezumabIgG4 (humanized mouse MABT5102)Aβ 13–24 (conformational epitopes) Binds fibrillar, oligomeric, and monomeric AβPhase III up to 60 mg/kg SC (every 2 weeks) for at least 260 weeks[389,390]
BAN2401IgG1 (humanized mAb158)recognizes Aβ protofibrilsPhase I: 2.5, 5 and 10 mg/kg
Phase II: 2.5, 5 and 10 mg/kg
[391,392,393]
Aducanumabb IgG1 (BIIB037/BART full human)recognizes Aβ oligomer and fibrils [394,395]
Table 3. Summary of pharmacological agents enhancing autophagy.
Table 3. Summary of pharmacological agents enhancing autophagy.
Drug NameTypePathologyModel TypeResultsReferences
1-methyl-4-phenylpyridiniumDopaminergic neurotoxinCulture models of Parkinson’sMouseinduce buildup of autophagic vacuoles[218,219]
Rapamycinselective inhibitor of TORC1Alzheimer’s DiseaseMouseameliorates Aβ and tau in AD mouse models[226,228]
Latrepirdinethe stimulator of Atg5-dependent autophagyAlzheimer’s DiseaseMousereduces Aβ in mouse models[220]
MetforminProtein phosphatase 2A agonistAlzheimer’s DiseaseClinical TrialsInhibits tau hyperphosphorylation in AD clinical trials[220,230]
LithiumUlK1 Kinase activatorAlzheimer’s DiseaseExperimental/Clinical trialsAMPK activation and induces autophagic activation[222]

Share and Cite

MDPI and ACS Style

Hussain, R.; Zubair, H.; Pursell, S.; Shahab, M. Neurodegenerative Diseases: Regenerative Mechanisms and Novel Therapeutic Approaches. Brain Sci. 2018, 8, 177. https://doi.org/10.3390/brainsci8090177

AMA Style

Hussain R, Zubair H, Pursell S, Shahab M. Neurodegenerative Diseases: Regenerative Mechanisms and Novel Therapeutic Approaches. Brain Sciences. 2018; 8(9):177. https://doi.org/10.3390/brainsci8090177

Chicago/Turabian Style

Hussain, Rashad, Hira Zubair, Sarah Pursell, and Muhammad Shahab. 2018. "Neurodegenerative Diseases: Regenerative Mechanisms and Novel Therapeutic Approaches" Brain Sciences 8, no. 9: 177. https://doi.org/10.3390/brainsci8090177

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

Article Metrics

Back to TopTop