Next Article in Journal
Inferring Single-Cell 3D Chromosomal Structures Based on the Lennard-Jones Potential
Next Article in Special Issue
Glycogen Synthase Kinase 3 Beta Regulates the Human Aryl Hydrocarbon Receptor Cellular Content and Activity
Previous Article in Journal
Metabolic Dysfunction in Spinal Muscular Atrophy
Previous Article in Special Issue
Dual Actions of A2A and A3 Adenosine Receptor Ligand Prevents Obstruction-Induced Kidney Fibrosis in Mice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

When Good Kinases Go Rogue: GSK3, p38 MAPK and CDKs as Therapeutic Targets for Alzheimer’s and Huntington’s Disease

by
Santosh R. D’Mello
Neugeneron, Dallas, TX 75243, USA
Int. J. Mol. Sci. 2021, 22(11), 5911; https://doi.org/10.3390/ijms22115911
Submission received: 10 May 2021 / Revised: 26 May 2021 / Accepted: 28 May 2021 / Published: 31 May 2021
(This article belongs to the Collection Feature Papers in Molecular Pharmacology)

Abstract

:
Alzheimer’s disease (AD) is a mostly sporadic brain disorder characterized by cognitive decline resulting from selective neurodegeneration in the hippocampus and cerebral cortex whereas Huntington’s disease (HD) is a monogenic inherited disorder characterized by motor abnormalities and psychiatric disturbances resulting from selective neurodegeneration in the striatum. Although there have been numerous clinical trials for these diseases, they have been unsuccessful. Research conducted over the past three decades by a large number of laboratories has demonstrated that abnormal actions of common kinases play a key role in the pathogenesis of both AD and HD as well as several other neurodegenerative diseases. Prominent among these kinases are glycogen synthase kinase (GSK3), p38 mitogen-activated protein kinase (MAPK) and some of the cyclin-dependent kinases (CDKs). After a brief summary of the molecular and cell biology of AD and HD this review covers what is known about the role of these three groups of kinases in the brain and in the pathogenesis of the two neurodegenerative disorders. The potential of targeting GSK3, p38 MAPK and CDKS as effective therapeutics is also discussed as is a brief discussion on the utilization of recently developed drugs that simultaneously target two or all three of these groups of kinases. Multi-kinase inhibitors either by themselves or in combination with strategies currently being used such as immunotherapy or secretase inhibitors for AD and knockdown for HD could represent a more effective therapeutic approach for these fatal neurodegenerative diseases.

1. Introduction

Alzheimer’s disease (AD) is a mostly sporadic brain disorder characterized by progressive cognitive decline resulting from neurodegeneration that starts in the entorhinal cortex and progresses to the hippocampus and large portions of the cerebral cortex [1,2,3,4]. In contrast, Huntington’s disease is a monogenic inherited disorder characterized by progressive motor deficits and psychiatric disturbances resulting from neurodegeneration largely localized to the striatum and to a lesser extent specific neuronal populations in the cerebral cortex [5,6,7]. It is generally believed that neurodegeneration in AD is caused by elevated levels of an abnormal form of the amyloid-β (Aβ) peptide, Aβ42, that forms extracellular oligomers and aggregates, and the hyper-phosphorylation of the microtubule-associated protein Tau, promoting its disassociation from axonal microtubules and deposition in insoluble neurofibrillary fibrillary tangles [1,2,3,4]. Oligomeric forms of both Aβ42 and Tau affect synaptic function and neuronal survival. Additionally, glial cells also become dysfunctional contributing to disease pathogenesis. HD is caused by the abnormal expansion of a polyglutamine (polyQ) stretch within the N-terminal region of the huntingtin (Htt) protein resulting in its misfolding. Although how the production of mutant polyQ-expanded Htt (or mut-Htt) causes neurodegeneration is unresolved, the consensus view is that the mut-Htt protein is abnormally phosphorylated, proteolytically cleaved and aggregates of the N-terminus fragment of the cleaved protein accumulate in the nucleus disrupting transcription. Aggregates of mut-Htt are also found in the cytoplasm where they are disrupt a variety of cellular processes [5,6,7]. Although with different etiologies (largely sporadic versus strictly genetic) and affecting largely non-overlapping neuronal populations resulting in distinct clinical features, the two diseases (and many other age-associated neurodegenerative diseases) share a variety of molecular and cellular commonalities. In addition to abnormal protein aggregation, these include mitochondrial dysfunction, elevated oxidative stress, endoplasmic reticulum (ER) stress, deregulation of autophagy, and abnormal post-translational modifications resulting from deregulated activity and functioning of the enzymes that mediate these modifications that in turn impact the functioning of other enzymes, macromolecules and cellular processes not just in neurons, but glial cells also. Additionally, aberrant activation of astrocytes and microglia induce the release of toxic cytokines that injure neurons. Among the enzymes that have been most convincingly implicated in the pathogenesis of neurodegenerative diseases are protein kinases, a family of about 500 proteins, most of which fall into two types-serine/threonine kinases, which phosphorylate serines and/or threonines in their target protein, and the tyrosine kinases. Among the hundreds of serine-threonine protein kinases, only a small number have been seriously implicated in the pathogenesis of neurodegenerative diseases and within this group, less than a dozen common kinases have been described to play a key role in diverse neurodegenerative disease with distinct etiologies. Prominent among these are glycogen synthase kinases (GSKs), the mitogen-activated protein kinases (MAPKs) and the cyclin-dependent kinases (CDKs). This review focuses on these three kinases and their contributions to the pathogenesis of AD and HD. The review starts with a brief overview of AD and HD, focusing on the cellular and molecular abnormalities. A summary of the properties of each of the three kinases is then described before reviewing evidence for their involvement in each of the two neurodegenerative disease, and their merits as targets for the development of therapeutics for the diseases. Although much of the currently-developed therapeutic approaches for AD target Aβ or Tau [8,9] and mut-Htt for HD [10], this review aims to impress upon the reader that the neuropathological action of these proteins requires or is regulated by the three kinases that the review focuses on and makes the case for targeting these kinases, possibly in combination with approaches that are currently being clinically tested, such as genetic knockdown or immunotherapy.

2. The Diseases

2.1. Alzheimer’s Disease

AD is a predominantly sporadic disorder with several risk factors the best defined of which are aging as well as mutations and variations of a large number of susceptibility genes, among which the most studied and accepted are expression of the ApoE4 (apolipoprotein E4) isoform and mutations in TREM2 (triggering receptor expressed on myeloid cells-2), a protein expressed within the brain by microglia [1,2,3,4]. For a comprehensive listing and description of the susceptibility genes for AD the reader is referred to a review by Verheijen and Sleegers [11]. In ~4% of the total cases, however, AD is familial, caused by mutations in genes including those encoding Aβ precursor protein (APP), presenilin-1 (PS1) and presenilin-2 (PS2) [1,2]. APP is the protein from which Aβ peptide is derived from by proteolysis while PS1 or PS2 are the catalytic components of the γ-secretase complex, which along with b-secretase/BACE1, cleaves APP to produce Aβ. Altered activities of BACE1 and γ-secretase in AD results in increased production of Aβ. This along with reduced proteolytic degradation of Aβ extracellularly by proteases, such as insulin-degrading enzyme (IDE) and neprilysin, results in its accumulation as neurotoxic Aβ oligomers and fibrils [12,13]. Amyloid plaques, a neuropathological hallmark of AD composed of insoluble deposits of Aβ aggregates are widely believed to represent sites at which oligomeric Aβ species are sequestered as a protective mechanism against their toxicity [14]. Hyperphosphorylation of Tau results in its disassociation from axonal microtubules leading to destabilization of microtubules, thus affecting axonal structure and transport [14,15]. Disassociated Tau forms neurotoxic fibrils and insoluble fibrillary tangles (NFTs) in the cytoplasm of neurons representing another neuropathological hall mark of AD [15]. Although the three AD-causing genetic mutations are linked to Aβ synthesis, NFTs are more closely correlated with cognitive impairments [15,16]. Furthermore, although neuronal death is the defining feature of the disease, it is synaptic degeneration and dysfunction that initiates cognitive impairment the clinical feature that best characterizes AD [17,18,19,20]. The accumulation of extracellular Aβ plaques triggers the activation of microglia and astrocytes that release inflammatory cytokines. Although with the purpose of eliminating the plaques, chronic release of inflammatory cytokines, a process referred to as neuroinflammation, promotes synapse elimination and neuronal death [21,22]. The AD susceptibility genes, ApoE and TREM2, respectively, contribute to the clearance of Aβ and reducing neuroinflammation, respectively, providing an explanation for why variant and mutant isoforms deficient in these functions increases the risk for AD [23,24,25]. Dysfunction of astrocytes and microglia also contribute to deregulation of glutamate homeostasis, resulting in the elevation of extracellular glutamate and the triggering of excitotoxicity in neurons through the overactivation of NMDA receptors, a subtype of ionotropic glutamate receptors [26]. Since astrocytes contribute to Aβ catabolism, astrocyte dysfunction could contribute to the pathogenic accumulation of Aβ [26]. Additionally, while astrocytes and microglia ultimately cause neuroinflammation, under normal circumstances these cells secrete cytokines that support the survival and functioning of neurons, a function which is lost when these cell types become functionally compromised [26,27]. Although less studied than astrocytes and microglia, emerging evidence suggests that impaired functioning of oligodendrocytes and their progenitor cells (which remain through adulthood) may also play a causal role in AD. For example, breakdown of myelin is enhanced in AD, occurs early in the diseases process and has been found to correlate with the spreading of NFTs [28,29,30,31], and both oligodendrocytes and oligodendrocyte progenitor cells (OPCs) phagocytose Aβ a function that is impacted in AD [32]. One recent study described that oligodendrocytes may facilitate the spread of pathogenic Tau in the brain even in the absence of neuronal Tau [33]. Taken together, although the overwhelming focus on Aβ and Tau pathology has led to a neuron-centric model for AD pathogenesis, findings from more recent research is making it increasingly clear that dysfunction of glial cell types is also an important contributor.
Another widely described pathogenic event in AD is the aberrant entry of neurons into the cell cycle leading to mitotic failure and death [34,35,36,37,38,39]. Although the significance of aberrant cell cycle entry to the triggering of AD is still to be fully resolved, it is considered by many to be, along with Aβ and Tau pathology, a third major cellular abnormality underlying neurodegeneration in AD (and several other neurodegenerative diseases). Several lines of evidence support this conclusion, including the activation of cell cycle promoting cyclin-dependent kinases (CDKs) and cyclins, misregulation of other cell cycle regulatory proteins, including retinoblastoma protein (Rb), CDC25 phosphatases, and the increase in aneuploidy and mitotic morphology in degenerating neurons and areas of the brains of AD mice and patients [35,36,40,41,42,43]. It has been suggested that aneuploidic neurons can survive in the brain for many years but selectively die during aging providing an explanation for the late-onset neurodegeneration [34,44]. Work in mice has revealed that increase in cell cycle markers precedes the elevation in Aβ and Tau hyperphosphorylation [45]. Consistent with a causal role for cell cycle entry, CDK inhibitors block neuronal death in cell culture models of AD (Rao et al., 2020; Herrup 2012).
Although the precise mechanisms and extents of contribution remain unclear, it is widely accepted that oligomeric Aβ and hyperphosphorylated Tau fibrils along with excitotoxicity, neuroinflammation and abortive cell cycle entry and a variety of other cellular disturbances, including impaired cholinergic signaling [46], ER stress [47,48], disrupted autophagic clearance of protein aggregates [49,50,51,52,53], mitochondrial dysfunction [54], oxidative stress [55], and deregulation of iron metabolism [56] combine to cause progressive synaptic failure and neuronal death in the AD brain.
There are no effective disease-modifying drugs for AD with existing medications, primarily acetylcholine esterase inhibitors and the NMDA receptor antagonist, memantine, serving to modestly relieve symptoms [57]. Clinical trials conducted so far aimed at developing disease-modifying therapies have all been unsuccessful. Recent studies have concluded that molecular changes in the AD brain begin decades before neuropathology and behavioral deficits are displayed by patients indicating that therapeutic intervention must be delivered early in the disease process for effectiveness [58,59]. In this regard it is noteworthy that much of the recent efforts at developing disease-modifying therapies for AD have focused on Aβ and Tau and include immunotherapy, drugs to reduce Aβ production or increase clearance, and drugs to inhibit Tau phosphorylation or polymerization, but these have yet to show success in the clinic [57,60,61]. Given the evidence that abnormal activation of kinases, such as GSK3, p38 MAP kinase and CDKs is an early and critical event in the disease process [62,63] that precedes Aβ and Tau oligomerization/fibrillization and actually contribute to it, stronger emphasis must be placed on the development of kinase inhibitors that may be utilized along with Aβ and Tau-targeted therapies.

2.2. Huntington’s Disease (HD)

HD is an autosomal dominant, progressive and fatal neurodegenerative disease caused by the abnormal expansion of a CAG trinucleotide repeat located in exon 1 of the Htt gene resulting in the expansion of a polyglutamine stretch in the N-terminus region of the Htt protein, rendering the mutant protein prone to misfolding [5,6,7]. Wild-type Htt is a large protein (~350 kD) present in both the cytoplasm and nucleus where it interacts with a large number of proteins [64]. Htt plays a critical role during embryogenesis and, consistent with the large repertoire of interacting-proteins, is involved in numerous cellular processes postnatally including regulation of transcription [65], nucleocytoplasmic trafficking [66], axonal transport [67], DNA repair [68], autophagy [69], mitophagy [70] and cell division [71]. Although mut-Htt (like wild-type Htt) is expressed ubiquitously, HD is characterized by selective degeneration of medium-spiny neurons (MSNs) of the striatum and, to a lesser extent, pyramidal neurons in specific layers of the cortex [5]. Mut-Htt forms oligomers and large aggregates in the nucleus and cytoplasm that disrupt neuronal function and promote degeneration in vulnerable brain regions. Although not well studied, neuronal loss and pathology has also been described in the hippocampus, thalamus and cerebellum in HD [72]. Convincing evidence indicates that mut-Htt oligomers are the toxic species whereas the aggregates that form inclusions may be protective [73,74,75]. It is generally recognized that both loss of normal function and acquisition of a toxic function by the mutant protein are involved in HD pathogenesis although the underlying molecular mechanisms remain largely unresolved. Dysregulation in neurotransmitter systems, including the glutaminergic, GABAergic and dopaminergic systems have been documented in HD mouse models and patients [76,77]. A role for mitochondrial dysfunction and oxidative stress has been amply documented [5,78,79]. A major cause of synaptic and neuronal loss in HD is excitotoxicity resulting from glutamate dyshomeostasis [77]. Chronic activation of microglia and astrocytes and the presence of neuroinflammation is a consistent feature of HD [80,81,82]. It is known that mut-Htt promotes release of inflammatory cytokines from astrocytes and microglia [83,84]. As observed in AD and several other neurodegenerative diseases, abnormal immune system activation has been documented in HD patients and mice leading to the elevation of IL-6 and other inflammatory cytokines and this has been reported to occur even before the appearance of diseases symptoms suggesting that immune system dysfunction contributes to brain pathology [84,85]. Interestingly, the cytokines released in HD are fewer and more localized within the striatum and other brain regions than the more general pattern seen in AD and other neurodegenerative diseases [85]. Whether neuroinflammation triggers development of neuropathology and behavioral deficits in HD has yet to be firmly established, there is wide consensus that neuronal death during the diseases provokes an neuroinflammatory response that exacerbates neurodegeneration [5,82].
While clearly a consequence of polyQ-expansion, other regions within the Htt protein and post-translational modifications within them are critically involved in regulating disease pathogenesis. Most studied in this regard is the N-terminus region of the protein containing the polyQ expansion. The highly-conserved 17 amino acids before the polyQ stretch, referred to as N17, which targets Htt to different subcellular compartments and regulates its folding, protein-protein interactions and clearance, is phosphorylated at three residues Thr3, Ser13 and Ser16 [86,87,88]. Phosphorylation at Ser13 and Ser16 is reduced in polyQ-expanded Htt whereas replacement of these two residues with phosphomimetic residues inhibits aggregation and neurotoxicity by mut-Htt both in tissue culture and in mice [86,87]. Initial studies pointed to IKKβ as being the kinase that phosphorylates Ser13 and Ser17 [86] although a subsequent study reported that the phosphorylating kinase is Casein Kinase-2 (CK2) [89]. Similarly, phosphorylation of Thr3 by IKK reduces aggregation and neurotoxicity of mut-Htt in cultured cells and in Drosophila [88,90]. Additionally, deregulations in signaling pathways cause disease-associated posttranslational modifications in other proteins within neurons, including DARP32 (dopamine and cAMP-regulated phosphoprotein), Tau, the CREB transcription factor, the mitochondria fission regulating protein Drp1 (dynamin-related protein 1), and several others [91]. The deregulations in many signaling pathways occur prior to mut-Htt aggregation and some evidence indicates that they play a key role in promoting mut-Htt aggregation and disease pathogenesis [91].
An accumulating body of evidence supports a role for Tau dysregulation in HD disease pathogenesis. Tau phosphorylation is increased in vulnerable regions of the brains of HD mice and patients where it accumulates in aggregates [92,93,94,95,96]. Additionally, and as a result of splicing alterations, the expression levels of specific Tau isoforms are changed in HD brain tissue causing disruption of the nuclear membrane [97,98]. Based on the documented Tau pathology in HD several investigators have referred to HD as a tauopathy, placing it in the same disease category as AD [99,100,101,102].
Although a monogenic disorder caused by a single mutation (trinucleotide expansion) in a gene identified about 30 years ago and the availability of a variety of invertebrate and rodent models, HD remains without disease-modifying therapies. Clinical trials that have targeted intracellular alterations such as mitochondrial dysfunction, reactive oxygen species (ROS) accumulation and oxidative damage, neurotransmitter systems, and pathways that modify mut-Htt modifications and aggregation has so far been ineffective. Some studies have described neuroprotection by histone deacetylase-3 (HDAC3)-selective inhibitors, both in cell culture and mouse models of HD [103,104,105,106,107]. Based on the finding that knockdown of Htt in adult HD mice has no obvious adverse effects, much of the current therapeutic focus has been on approaches that lower Htt/mut-Htt expression [108,109]. These approaches include ones targeting expression at the mRNA level, including antisense oligonucleotides (ASOs) and RNA interference or at the DNA level, including CRISPR-Cas9, transcription activator-like effector nuclease (TALEN) and zing-finger proteins (ZNFs) [108,109]. Based on bioinformatic analyses, allele-selective ASOs have recently been developed that knockdown mut-Htt allele selectively. Although a promising avenue, the issue of off-target effects of the ASOs and other Htt-knockdown approaches in humans is a concern. A recent report described the identification of a chemical that binds expanded CAG tracts and promotes contraction of the repeats, which could have value in the treatment of HD and other CAG-repeat expansion disorders [110,111].

3. The Enzymes

3.1. (A) GSK3

GSKs, a family of three serine-threonine protein kinases—GSK1, GSK2 and GSK3—were identified by their ability to phosphorylate glycogen synthase, the rate limiting enzyme of glycogen metabolism [112,113,114]. GSK3 is the best studied of the GSKs, which is known to regulate a variety of cellular functions besides glucose metabolism. As in much of the literature, for the rest of this review GSK refers to the GSK3 protein. Two paralogs, commonly referred to as isoforms, of GSK3 are produced from separate genes: GSK-3α and GSK3β with molecular weights of 51 kDa and 47 kDa, respectively [115]. The two isoforms share ~95% amino acid identity and are therefore thought to phosphorylate many of the same proteins. Although involved in common functions, GSK3α and GSK3β have non-overlapping functions best exemplified by the finding that knockout of GSK3β in mice generally results in embryonic lethality whereas global knockout of GSK3α results in viable animals although these mice develop age-related pathologies and a slightly shortened lifespan [116,117,118]. Specific non-overlapping functions for the two GSK isoforms in the brain have been described [119,120,121,122]. GSK3 is unusual among protein kinases in that it is generally constitutively-active making its inhibition, rather than activation, the primary mode of its regulation [123]. Alterations in GSK3 activity depends on the extent of phosphorylation of Ser-9 in GSK3β (or Ser-21 in GSK3α) which inhibits constitutive activity, and the phosphorylation at Tyr216 of GSK3β (Tyr279 of GSK3α) which restores and increases enzyme activity. Phosphorylation at other residues also contributes to increased GSK3 activity. In addition, GSK3 activity and function depend on its subcellular localization, the subcellular localization of its substrates, the level of its substrates, and its inclusion in multi-protein complexes that can control access to its substrates [123]. It has been reported that the activity of GSK3 can be increased through cleavage by calpains [124,125,126,127] and matrix metalloproteinase-2 [128] and inhibited by mono-ADP-ribosylation [129,130], acetylation [131,132] and citrullination [133]. Although largely a cytosolic protein, GSK3 is present in the nucleus and mitochondria. In the nucleus, phosphorylation by GSK3 modulates the activities and functions of many transcriptional regulators, many of which are known to promote neurodegeneration, including c-jun [134,135], HDAC3 (histone deacetylase-3) [136], HDAC4 [137] and DNMT1 (DNA methyltransferase-1) [138]. In contrast to most other kinases, the substrates of GSK3 generally (but not always) need to phosphorylated by another kinase before they are phosphorylated by GSK3 [123]. Despite this requirement, GSK3 is believed to phosphorylate more substrates than any other protein kinase with over 100 substrates being identified in biochemical studies and a much larger number based on computational analyses [139,140].

3.1.1. GSK3 in the Brain

GSK3α and GSK3β are differently-regulated during brain development with GSK3β being highly expressed during neurogenesis while GSK3α is poorly expressed during that period. In the mature brain GSK3β is widely present whereas GSK3α shows a more restricted pattern of expression with highest expression in in the cerebral cortex, striatum hippocampus, and Purkinje cells of the cerebellum [141]. Likely because of it abundance and widespread expression in the brain and the prenatal lethality of GSK-3α knockout mice, research on the role of GSK3 in the brain have until recently focused almost exclusively on GSK3β. Multiple roles have been proposed for GSK3β in the adult brain including the negative regulation of neurogenesis in the hippocampus [142,143,144], regulation of synaptic plasticity, learning and memory [145,146,147,148], and stimulation of the inflammatory function of microglia [149,150,151]. While with multiple beneficial functions, elevated GSK3 activity promotes death of neurons in culture [152,153,154] and in the brains of mice [155].

3.1.2. Roles of GSK3 in AD

GSK activity is upregulated in the hippocampus of AD patients and is associated with phosphorylated Tau and NFTs [156,157]. Several studies have established that Tau phosphorylates GSK3 at multiple disease-relevant sites in mouse models of AD and in the patient brains, and is a central kinase in Tau hyperphosphorylation causing disassociation of Tau from microtubules and fibrillization [15,158,159,160,161] (Figure 1A). GSK3 has also been reported to stimulate Aβ production whereas its pharmacological inhibition or knockdown reduces processing of APP to Aβ [161,162,163,164] (Figure 1A). Consistently, GSK3 over-activation can increase the activity of PS1 [165,166,167] and the expression of BACE1 [168,169]. Elevated Aβ activates GSK, which then through increased processing of APP also produces more Aβ resulting in a positive-feedback loop. Aβ-mediated GSK3 activation also increases Tau phosphorylation and induces its fibrillization linking Aβ to Tau dysfunction via GSK3 [170,171]. Inhibition of GSK3 reduces Aβ levels, tau hyperphosphorylation, and cognitive deficits in mice [15,158] (Figure 1A).
As described above, an important cellular mechanism that is impaired in AD is autophagy, a process that degrades and clears misfolded and aggregated proteins as well as other dysfunctional macromolecules and organelles. In the context of AD, impaired autophagy is thought to be responsible in major part for the accumulation of Aβ and Tau aggregates [50,52]. Not surprising, drugs that activate autophagy have been found to be neuroprotective in cell culture and mouse models of AD [172,219,220,221]. GSK3 is a well-established inhibitor of autophagy and some studies have shown that inhibitors of GSK3 reduce AD pathology by stimulating autophagy [172,173].
In addition to directly promoting neuronal death in neurons, GSK3 promotes neuroinflammation in AD though the regulation of the expression and release of harmful cytokines from glial cells [147,149,151,174]. Adult neurogenesis, which serve to replace neurons, is sharply reduced in the AD hippocampus exacerbating the loss of hippocampal function [143]. This reduction of neurogenesis in AD has been found to be GSK3-dependent. Inhibiting GSK3 promotes neurogenesis and reduces neuronal death in the adult hippocampus [142]. In sum, convincing results indicate that GSK3 activation occurs early in the disease process and promotes all the major pathogenic changes that cause or contribute to AD pathogenesis [62,222,223,224] (Figure 1A).

3.1.3. GSK3 as a Therapeutic Target in AD

Several brain-penetrant GSK3 inhibitors have been evaluated in cell culture, slice and animal models of AD. A number of laboratories have tested lithium, a drug commonly used in the treatment of bipolar mood depression but known to inhibit GSK3 [225,226]. Lithium administration inhibits GSK3 in mouse models of AD and promotes Aβ and Tau neuropathology, facilitates LTP (long-term potentiation) induction and improves cognitive performance [175,227,228,229]. Lithium also reduces neuronal dysfunction in an Drosophila model of AD in which Aβ is inducibly-overexpressed in adult flies [230]. However, lithium is a poor drug candidate for use in humans because of its narrow therapeutic window and serious side effects, including neurotoxicity, particularly in the elderly [231,232,233]. Lithium is also known to inhibit other enzymes and that its beneficial effects in AD mice is due to inhibition of GSK3 has yet to be established [234,235]. Despite its limitations, lithium has been tested in patients with mild cognitive impairment (MCI) and AD with confusing results. While a couple of studies found slight improvement in cognitive function [236,237], others studies reported no improvement or even an increase in dementia [238,239,240]. Tideglusib (also referred to as NP12 or NP031112) is another selective GSK3 inhibitor that reduces Tau phosphorylation, decreases Aβ deposition, inhibits plaque-associated gliosis, protects neurons in the entorhinal cortex and hippocampus against cell death, and reduces memory deficits in AD mice [176]. Although tideglusib was well-tolerated in AD patients, it displayed no clinical benefit in an initial safety study and a Phase-II trial [241,242]. A GSK3 inhibitor developed by AstraZeneca, AZD1080, reduces Tau hyperphosphorylation in rats and synaptic plasticity deficits in rodents but was abandoned in Phase II trials because of severe side effects [243,244]. Given the impressive effects of GSK3 inhibitors in a variety of in cell culture and preclinical studies, the lack of efficacy in AD patients is surprising. It is likely that treatment in patients requires initiation of treatment earlier in the disease process. Alternatively, it is possible that the complex mechanisms of AD pathogenesis in humans will require the simultaneous targeting of multiple molecules, including but not limited to GSK3. The strategy of combinatorial drug therapy is well-accepted in the treatment of cancer [245,246,247].
Another highly potent GSK3 inhibitor developed by AstraZeneca belonging to the pyrazine class and with excellent drug-properties, AZD2858, was effective at reducing Tau phosphorylation and gliosis in the hippocampus but did not proceed to clinical trials because of failure to pass preclinical toxicology studies [243]. A recent study described that a GSK inhibitor developed by Sanofi, SAR502250, reduced Tau hyperphosphorylation in the cortex and spinal cord in P301L human transgenic mice [248]. Additionally, this compound protected against neuronal loss resulting from Aβ treatment and reduced cognitive impairment in two separate mouse models of AD, a transgenic mouse model and the Aβ-infusion model [248]. The effect of SAR502250 on Aβ or Tau neuropathology or on neuroinflammation was not examined in this study and therefore how it exerts its beneficial effects have yet to be determined. A thiazole GSK3 inhibitor, AR-A014418, inhibits tau phosphorylation and is neuroprotective in cell culture and inhibited Aβ neurotoxicity in hippocampal slices [249]. AR-A014418 also reverses axonal transport defects and behavioral deficits in Tau-overexpressing Drosophila [250]. However, another study conducted in young rats reported that AR-A014418 did not inhibit Tau phosphorylation [251]. Chronic low-dose administration of another GSK3 inhibitor, AM404, reduced Aβ production, tau hyperphosphorylation, neuroinflammation and cognitive impairment in AD mice [244,252]. However, at higher doses this compound had detrimental effects on brain function [253].
Among other GSK3 inhibitors that have been tested to a limited extent in AD-related models are a set of isonicotinamides, which reduce Tau activity in triple-transgenic AD mice [254]. In another study several GSK3 inhibitors belonging to the aminopyrimidine, indurubin, alsterpaullone, thiazole classes were tested for their ability to suppress phosphorylation of Tau at Ser-396 in normal rats, increased phosphorylation of which is associated with AD [251]. It was found that CHIR98014, an aminopyrimidine, reduced Tau phosphorylation in the cortex and hippocampus, while two other GSK3-inhibiting compounds, an alsterpaullone and SB216763, reduced phosphorylation only in the hippocampus suggesting region-specific regulatory mechanisms for Tau phosphorylation, including differences in the pattern of Tau phosphorylation or the involvement of another kinase or phosphatase in one brain region and not another. Another study described that SB216763, an ATP competitive-inhibitor of GSK3, protected against neuronal damage resulting from intracerebroventricular infusion of Aβ in mice, but had very modest effects on gliosis and behavioral deficits [255]. Surprisingly, in control mice SB216763 induced inflammation and behavioral deficits possibly due to inhibition of constitutive GSK3 activity [255]. Antisense oligonucleotides against GSK3β have also been tested in SAMP8 mice, which display accelerated aging along with increased Aβ levels, Tau hyperphosphorylation, neuroinflammation and cognitive deficits and have hence been used as a model of AD [177,256,257]. Knockdown of GSK3β reduced oxidative stress in the brain and improved learning and memory of SAMP8 mice [177].

3.1.4. Roles of GSK3 in HD

Although predominantly cytosolic, GSK3 accumulates and co-localizes in lipid rafts with mut-Htt in cultured neurons and in presymptomatic HD mice, an alteration that has been suggested to contribute to disease development [258]. While the striatum and the cortex are the most affected brain regions in HD, neuronal loss and dysfunction in the hippocampus has also been described [259,260,261,262] An upregulation of GSK3β expression (mRNA and protein) and activity has been observed in the hippocampus of both HD mice and patients and in [180]. In HD mice the increase in GSK3β induces Tau phosphorylation and caspase-3 activation in the dentate gyrus leading to neuronal death [180]. Other studies have also described GSK3-mediated Tau phosphorylation and aggregation in HD, raising the possibility of common pathogenic mechanisms in HD and AD, at least with relation to Tau [92]. In addition to its actions in neurons, increased GSK3β expression and activity is observed in astrocytes within the hippocampus where it promotes release of inflammatory cytokines [180]. Although the mechanism of increased GSK3 activity has not been studied, it is possible that it is a consequence of reduced activity of Akt in the HD brain [263]. It is well-established that phosphorylation of GSK3 by Akt inhibits its activity in vitro and in vivo [264,265].

3.1.5. GSK3 as a Therapeutic Target in HD

Pharmacological inhibitors of GSK3 or expression of a dominant-negative form of GSK3β are neuroprotective in cell culture models of HD [106,258,266]. An important target of GSK3 is HDAC3, a protein that is required for the neurotoxic effect of mut-Htt in cultured neurons [106,107] and promotes neurodegeneration and behavioral deficits in HD mice [103,104,105]. Inhibition of GSK3 using pharmacologically or through expression of a dominant-negative construct inhibits both HDAC3 and mut-Htt toxicity in cultured neurons [106]. In vivo protection by GSK3β was reported in a C. Elegan model of HD, although the inhibitor used in the study was lithium, which as described above, has other cellular targets [267]. Chronic treatment of HD mice with lithium produces variable effects with some mice displaying marked improvement in motor function whereas other mice do not [268]. Life span was not extended by lithium treatment in any of the mice. Another study described substantial benefit of lithium in HD mice only when it was co-administered with valproic acid, which itself has multiple targets, including the inhibition of HDACs [269,270,271]. A similar requirement was described in another study that reported additive protective effects of lithium and the mTOR inhibitor, rapamycin, in a Drosophila model of HD [272]. These results suggest that multiple signaling molecules and pathways may be involved in HD pathogenesis and full rescue would therefore require inhibiting multiple targets.
In contrast to the conclusions of the aforementioned studies, the results of a few other studies indicate that GSK3 plays a neuroprotective role in HD. One study described reduced GSK3 activity in the striatum of HD mice and patients [273]. This study showed that moderate overexpression of GSK3β in the striatum of HD mice attenuates brain atrophy, motor impairment and cognitive deficits [273]. One study described that elevation in glycogen synthase activity leading to increased glycogen synthesis results in enhanced autophagic flux which protects against mut-Htt toxicity in cell culture models [274]. Although this would suggest that an increase in GSK3 activity would be neuroprotective, whether this is the case was not examined in the study. Whether some of these differences with regard to toxicity versus neuroprotection are due to opposing actions of the two GSK3 isoforms and/or selectivity of antibodies used with regard to the two isoforms is unclear.

3.2. (B) p38 MAPK

Along with the two other families of mitogen-activated protein kinases (MAPKs), JNKS (c-jun N-terminal kinases) and ERKS (extracellular signal-regulated kinases), the p38 MAPKs are enzymes that regulate cellular responses to a wide range of extracellular signals [275,276]. p38 MAPKs were previously referred to as stress-activated protein kinases (SAPKs) because they are activated by various stimuli that are stressful or noxious to cells, including osmotic stress, oxidative stress, heat shock, genotoxic and DNA-damaging agents, and pathogen proteins [276,277]. In many cases p38 MAPK responds to such cellular stresses by mediating an inflammatory response [278]. Mammals express four p38 MAPKs that are all about 38 kDa is size and encoded by distinct genes: p38α (MAPK14), p38β (MAPK11) p38γ (MAPK12) and p38δ (MAPK13) [277,279]. Analyses of the primary sequences of the four isoforms reveals over 60% overall sequence homology and greater than 90% homology in the kinase domains. Most studies have focused on p38α and p38β. p38α is expressed ubiquitously and highly, whereas p38β is expressed at lower levels and in many tissues with highest expression in the brain. The two proteins share functions, including the mediation of inflammatory responses and regulation of cell proliferation, differentiation, survival and death [275,277]. However, global knockout of p38α in mice causes embryonic lethality whereas p38β knockout mice are phenotypically normal and fertile indicating unique functions of p38α that cannot be compensated for by the other isoforms [280,281]. In contrast to the widespread expression of p38α and p38β, p38γ and p38δ are expressed in a tissue-specific manner and have more specialized functions although what these are have not been fully identified. Knockout mice lacking either p38γ and p38δ or double-knockout mice lacking both isoforms have on obvious abnormalities [280,281].
p38 MAPKs are activated by phosphorylation within a Thr-Gly-Tyr motif in the activation loop by the dual-specificity MAPK kinases, MKK3 and MKK6 [277,282]. A large number of cytoplasmic and nuclear proteins have been identified as p38 MAPK substrates, including scaffold proteins, cytoskeletal proteins, signaling proteins, chaperones, and transcriptional factors and regulators. To activate nuclear substrates, activated p38 translocates to the nucleus by mechanisms that have not been fully resolved (p38 MAPKs lack nuclear import or nuclear export motifs). Following substrate phosphorylation p38 MAPKs are inactivated by dephosphorylation of the Thr-Gly-Tyr motif by dual-specificity phosphatases (DUSPs), also known as MAPK phosphatases (MKPs) located in the cytoplasm and nucleus. Dephosphorylation by MKPs in necessary for translocation of p38 MAPK out of the nucleus. Inactivation of p38 MAPK is also promoted by transcriptional feedback, reduced activity of upstream kinases and the termination of the activating stimuli [277,282,283]. In some cases p38 MAPKs can be activated through phosphorylation of non-canonical residues independent of MKK3 and MKK6 [284,285]. Despite the high level of structural similarity, the four 38 isoforms are differentially sensitive to pharmacological inhibitors. For example, while p38α and p38β are sensitive to pyridinyl imidazole inhibitors, p38γ and p38δ are insensitive to them [286].

3.2.1. p38 MAPK in the Brain

Most studies on the function of p38 MAPK in the brain have focused on p38α and p38β, which are expressed in neurons as well as astrocytes, microglia and oligodendrocytes [283,287]. Within the brain both isoforms are highly expressed in the cortex and hippocampus. Subcellular localization analysis of hippocampal neurons revealed that p38α is localized in all cellular compartments, including the nucleus, soma, neurites and synapses whereas p38β is predominantly nuclear. Most studies have not distinguished between p38α and p38β in determining their relative contributions to a biological action. During development, p38 MAPK regulates neuronal differentiation, neuronal migration, development of the neuronal skeleton and synapse formation [283,288]. In mature neurons p38 MAPKs regulates ion channel function, axonal transport and axonal regeneration [283]. p38 MAPKs has been shown to regulate learning and memory [289,290]. Several lines of evidence suggest that p38 MAPK, and specifically p38α, regulates synaptic plasticity by promoting synaptic depression and memory [291]. Activation of p38 MAPK contributes to LTP inhibition [292,293,294] and LTD (long-term depression) induction [295,296] in the hippocampus. Interestingly, conditional deletion of p38α in neurons or astrocytes of the mouse hippocampus revealed that deleting neuronal p38α has no effect on LTD whereas deletion in astrocytes abolishes it suggesting that the stimulation of LTD in neurons was driven by its action in astrocytes and establishing a role for p38α in astrocyte-neuron communication. Several studies have described that p38 MAPKs promotes neuronal death induced by a variety of noxious stimuli, including inflammatory cytokines and neurotoxins [297,298], oxidative stress [299,300], hypoxic insult [301], and excitotoxicity [302]. For example, following neuronal damage or injury to the brain or spinal cord, p38 MAPK (and primarily p38α) promotes chronic inflammation through release of cytokines, including IL-1b and TNFa, from astrocytes and microglia, which while designed to be helpful has an inhibitory effect on recovery from injury [181,287,303]. IL-1b and TNFa released by microglia through the action of p38 MAPK also inhibit LTP and therefore synaptic plasticity [304]. Through p38 MAPK, activated astrocytes and glia produce reactive oxygen species that exacerbate the injury to neurons thereby promoting neurodegeneration [181,287]. In addition to causing neuronal dysfunction and death through activation astrocytes and microglia, p38 MAPK can induce death of neurons cell autonomously by activating apoptotic signaling pathways [184,305].

3.2.2. p38 MAPK in AD

p38 MAPK activity is elevated in the hippocampus and cortex in both AD mice and patients relatively early in the disease process [306,307,308,309]. The increase is accompanied by elevated activity of MMK6 on of the two major activators of p38 MAPK [310]. p38 MAPK activation has been observed in both neurons and glial cells. In microglia and astrocytes p38 stimulates chronic release of inflammatory cytokines, which has been shown to be triggered, at least in part, by APP and Aβ that bind to cell surface receptors [181,182,183,198,311,312] (Figure 1B). Although initially helping in the clearance of Aβ, as described above, chronic release of inflammatory cytokines causes neuronal dysfunction and degeneration. A key target of p38 MAPK in its pro-neuroinflammatory role is MK2 (MAPK-activated protein kinase 2). Not surprisingly, MK2 activity is elevated in the brains of AD mice and its genetic deletion in cultured microglia prevents the release of inflammatory cytokines following Aβ treatment. [313]. Within neurons p38 MAPK phosphorylates Tau [187,188], activates pro-apoptotic signaling pathways [182,184,185,186] and promotes excitotoxicity [195,196,197] (Figure 1B). Although all four p38 proteins phosphorylate Tau, pharmacological experiments in mice have identified p38α as the isoform most responsible for Tau phosphorylation at pathogenic residues [189]. Inhibitors selective for p38α suppress neuroinflammation and protect against synaptic dysfunction, cognitive deficits and behavioral deficits in AD mice [189,314]. In these studies the effect of the inhibitors on p38β was not directly assessed, however, and given the structural and functional similarities between the two isoforms, contribution of p38β is possible and perhaps even likely. Some studies have shown that p38 MAPK, and primarily the p38α isoform, inhibits autophagy by phosphorylating and inhibiting the ULK1 complex, a protein complex that initiates the autophagic signaling pathway [199,200,315]. Thus, besides promoting other pathological mechanisms within neurons, p38 MAPK contributes to AD neuropathology by inhibiting autophagy.
Interestingly, and in contrast to p38α (and p38β), p38g has a protective effect in AD mice [316]. This protective effect involves phosphorylation of Tau by p38γ at Ser205 which results in disruption of postsynaptic excitotoxic protein signaling complexes activated by Aβ [316,317]. Neuronal deletion of p38γ in AD mice results in exacerbation of neural circuitry degeneration and cognitive defects as well as premature lethality, demonstrating a protective role for p38γ in AD [316]. Other studies have described that p38γ phosphorylates Tau at Thr50, which enhances the ability of Tau to promote microtubule assembly [187]. In contrast, p38δ phosphorylates Tau at sites that reduce the ability of Tau to promote microtubule assembly [318,319]. Taken together, these results indicate that p38α (and likely p38β and p38δ) promotes AD pathogenesis whereas p38γ exerts a protective role.

3.2.3. p38 MAPK as a Therapeutic Target in AD

Chemical inhibitors selective for p38α have been tested in Aβ and Tau mouse models of AD and found to be effective at reducing inflammatory cytokine production, Tau pathology, synaptic dysfunction and cognitive impairment [320]. Among these are MW01-2-069A-SRM [314,320], MW181 [189] and MW150 [192,321], and VX-745 [322]. Administration of NJK14047, a p38 MAPK inhibitory compound with demonstrated anti-inflammatory effects, was found to reduce microglial-induced neuroinflammation as well as Aβ deposition, neurodegeneration and memory impairment in the 5XFAD mouse model of AD [323]. As described above, besides promoting neurodegeneration by stimulating chronic inflammation, p38 MAP directly causes neuronal death by activating pro-apoptotic pathways and increasing levels of reactive oxygen species [182,184,186]. A study conducted in rats showed that the p38 MAPK inhibitor, PD169316, blocked pro-apoptotic signaling and reduced neuronal loss induced by intracerebroventricular injection of Aβ [324]. p38 MAPK inhibitors have been shown to attenuate Aβ-induced LTD impairment in hippocampal and entorhinal cortex slices [192,193,194]. Protection against cognitive impairment in AD was described with another p38 MAPK inhibitor possessing anti-inflammatory effect [325]. Another study described reduction of neuroinflammation and cognitive impairment through administration of a peptidic MK2 inhibitor, MMI-0100 [326]. As described above, MK2 is a downstream target of p38 MAPK.
Other p38 MAPK inhibitors have been reported to have beneficial effects in cell culture and rodent models of AD but the selectivity of these inhibitors for p38 MAPK has not been well-documented [327,328,329,330,331]. Some studies have used p38 MAPK inhibitors that are selective, but their ability to suppress neuropathology or improve behavioral performance in vivo was not been adequately evaluated [332].

3.2.4. p38 in HD

p38 MAPK activity is increased in the stratum of HD patients and in mouse models and this increase has been found to be associated with neuronal death [195,333,334,335]. One study attributed the increase in p38 MAPK activity to reduced activity of MKP-1, a phosphatase that inactivates p38 MAPK by dephosphorylation of the Thr-Gly-Tyr motif [336]. Pharmacological stimulation of MKP-1 reduces p38 MAPK activity and protects against the neurotoxic effect of mut-Htt expression in cultured cells and in mice with striatally-injected mut-Htt [336]. Neurodegeneration caused by mut-Htt is, at least in part, due to NMDA receptor-mediated excitotoxicity. Excitotoxic death in HD has been found to involve enhanced interaction between the NMDA receptor and the postsynaptic protein PSD95, which through abnormal activation of calpains results in the cleavage of STEP61 a phosphatase enriched in the striatum [337]. Whereas full-length STEP61 negatively regulates p38 MAPK in striatal neurons, calpain-cleaved STEP61 cannot dephosphorylate p38 MAPK and therefore unable to suppress excitotoxicity [337]. Calcineurin, is yet another phosphatase, that enhances mut-Htt neurotoxicity both in cell culture and in the striatum of HD mice by dephosphorylating at Ser421 of mut-Htt [338]. Pharmacological inhibition of calcineurin with FK506 results in elevated Ser421 phosphorylation, which protects against mut-Htt neurotoxicity [263,338,339]. SGK (serum-and glucocorticoid induced kinase) is a kinase that phosphorylates mut-Htt at Ser421. Somewhat counterintuitively, increased p38 MAPK activity in HD leads to induction of SGK activity in the striatum and cortex [334]. It is likely that the induction of SGK following p38 activation reflects a stress response in neurons degenerating due to p38 MAPK activation. p38 MAPK suppresses chymotrypsin-like protease activity leading to the accumulation and aggregation of mut-Htt [340]. Pharmacological inhibition of p38 MAPK leads to an increase in chymotrypsin-like protease activity and consequently enhanced clearance of mut-Htt.

3.2.5. p38 MAPK as a Therapeutic Target in HD

Studies using HD mice have found that the increased activation of p38 MAPK correlates with striatal degeneration indicating a causal role for p38 MAPK in HD-associated neurodegeneration [333,337]. The p38 MAPK inhibitor, SB-239063, protects striatal neurons cultured from HD mice from degeneration [195]. Similar neuroprotective effects of another p38 MAPK inhibitor, SB203580, was described both in HD mice and striatal cells overexpressing mut-Htt [341]. The mechanism by which p38 MAPK inhibition protects in mouse HD models is not fully clear but likely involves blockade of apoptotic signaling within neurons as well as inhibition of glia-mediated neuroinflammation. Other mechanisms by which p38 MAPK inhibitors could protect is through inhibition of calcineurin and activation of chymotrypsin-like protease activity, as described above.

3.3. (C) Cyclin-Dependent Kinases

The eucaryotic cell cycle is comprised of four phases – G1, S, G2 (phases during which the cell grows and duplicates its DNA) and M phase when mitosis occurs. Quiescent cells that lack sufficient nutrients to traverse the cell cycle reside in the G0 phase, a phase in which neurons and other postmitotic cells also reside. Two check-points, G1/S and G2/M, ensure that the cell is ready for DNA replication and that DNA replication is accurately completed, respectively. Progression through the various phases of the cell cycle is driven in large part by cyclin-dependent kinases (CDKs), a family of serine-threonine kinases. Activation of CDKs require heteromerization with specific cyclins, such that each CDK has one (or sometimes two) cognate cyclin protein interacting with and activating it [342,343,344,345]. Cyclin binding, which involves a motif called the cyclin box, changes the conformation of the CDKs to expose the substrate-binding region of the kinase [345,346]. The levels of most cyclins oscillate within the cell such that they are synthesized as the cell cycle enters the phase in which the cyclin is needed and rapidly degraded on exiting it which inactivates its CDK. Whereas many of the approximately 21 CDKs expressed in mammals play various roles including the regulation of transcription, cell cycle progression is under the charge of a core group of CDKs, including CDK1, 2, 3, 4 and 6.
The transition from G0 to G1 requires the activity of CDK3/cyclin C whereas CDK4-cyclin D and CDK6-cyclin D promote passage through G1. A key substrate of both CDK4 and CDK6 is the retinoblastoma protein (Rb) which sequesters the transcription factor E2F1 by interacting with it. Hyperphosphorylation of Rb by CDK4 and CDK6 during the G1 phase causes its disassociation from E2F, allowing E2F1 to bind to the promoters and induce expression of cell cycle promoting genes, including the cyclin A and E genes. The CDK2-cyclin E/A complexes promotes transition through S-phase followed by CDK1/Cyclin A- and CDK1/Cyclin B-mediated passage through G2, M-phase and cytokinesis. The activity of the CDK-cyclin complexes are inhibited by CDK-inhibitory proteins (CKIs) which act at different phases to slow down or prevent cell cycle progression. A multitude of other protein, including transcription factors and various inhibitory and activating kinases and phosphatases, modulate the activity of CDK-cyclins on cell cycle progression [343,344].
Neurons are kept in G0 through the constant inhibition of cell cycle-promoting CDKs along with activation of CDK inhibitors. However, aberrant activation of CDKs, often resulting from the loss of inhibitory mechanisms, allows neurons to enter the cell cycle leading to their death by apoptosis [34,347,348,349]. Increasing E2F1 levels by ectopic expression (which is normally caused by CDK4 and CDK6) induces apoptosis in neurons whereas suppression of E2F1 expression is neuroprotective in cultured neurons exposed to a variety of death-inducing stimuli and in animal models of neurodegeneration [350,351,352,353,354,355]. Interestingly, E2F1 is phosphorylated both in culture and in vivo by GSK3 and p38 MAPK [178,190,191]. Indeed, phosphorylation by GSK3 is necessary for E2F1-induced neuronal death [179]. In addition to E2F1, pharmacological inhibition of GSK3 also inhibits of expression of cyclin D, cyclin E, Rb phosphorylation and the activity of E2F1 [356]. These results indicate cross-talk between the cell cycle machinery and signaling by GSK3 and p38 MAPK.
An unconventional and somewhat unique member of the CDK family is CDK5, a widely-expressed protein that is not activated by cyclins and does not directly participate in cell cycle progression. CDK5 is activated through interaction with either of two activating proteins, p35 and p39, that are expressed only in the nervous system and predominantly in postmitotic neurons [357]. Because of the restricted expression pattern of its activating proteins, the kinase activity of CDK5 is restricted to the CNS where it is involved in a variety of critical functions, particularly during brain development but also during adulthood [357].

3.4. Role of CDKs in the Brain

Development of the mammalian brain requires rapid proliferation of neuroepithelial cells and their derivatives, neural progenitor cells (NPCs), which results in increased brain mass [358,359,360]. After a period of proliferation, increasing numbers of NPCs exit the cell cycle and start differentiating into neurons and other brain cell types [358,359,360]. Careful regulation of CDKs and other cell cycle components is critical for proper development and functioning of the brain. Deregulation of CDKs, their cyclin partners, or their regulators affects not only the generation of the proper number of NPCs but also the timing and extent of their differentiation to generate the proper number of postmitotic neurons and glial cells. Indeed, disruption of the balance between NPC production and cell cycle exit impacts neuronal differentiation and migration, and underlies a variety of neurodevelopmental disorders [361,362,363]. While primarily involved in promoting NPC proliferation, the presence of CDKs 1, 2 and 4 and their cognate cyclins are detectable in dendrites and axons where they inhibit neuronal maturation [364,365,366]. Abnormal increases in the activity of these CDKs result in aberrant phosphorylation of proteins in axons and dendrites, including Tau [367]. In contrast to the cell cycle-promoting CDKs, CDK5 promotes differentiation, maturation and migration of postmitotic neurons on the developing brain [360]. In the adult brain CDK5 plays a key role in synaptic plasticity, learning and memory [368,369]. CDK5 also prevents neurons from entering the cell cycle through interaction with E2F1. However, and as described below, overactivation of CDK5 can negatively brain function and can promote neurodegeneration [370,371,372]. One study has described that normally CDK5 activity is prevented from becoming overactive by cyclin E through direct binding [373]. Genetic ablation of cyclin E, which is expressed at high levels in terminally differentiated neurons, results in abnormal activation of CDK5 reduced number and volume of dendritic spines and impaired synaptic plasticity and memory [373].

3.4.1. CDKs in AD

A compelling body of evident implicate aberrant CDK activation and cell cycle entry as being causally involved in AD pathogenesis. Death of cultured cortical and hippocampal neurons resulting from Aβ treatment is preceded by induction of various cyclins and activation of CDKs and can be prevented by treatment with broad spectrum CDK inhibitors [374,375,376,377]. Increased activity of specific CDKs and/or cyclins or deregulation of other cell cycle-modulating proteins has also been described in vulnerable regions of the brain in both AD mice and patients [34,35,36,201,202,203]. In cycling cells expressing Tau, cell cycle progression coincides with Tau hyperphosphorylation and altered microtubule stability [378,379]. Similarly, in the AD brain activation of the cell cycle is believed to contribute to Tau hyperphosphorylation and aggregation as well as microtubule destabilization [201,380]. However, one study described that Aβ-induced stimulation of cell cycle markers requires Tau as the increase is not seen in Tau-deficient neurons [381]. Taken together, these findings suggest the possibility of a feed-forward loop in which deregulated cell cycle components cause Tau phosphorylation, which in turn stimulates the cell cycle. Intracerebral infusion of Aβ in mice results in increased expression of mitotic proteins prior to memory deficits, both of which are prevented by co-administration of the CDK inhibitor, flavopiridol, suggesting CDKs involvement in AD-related memory impairment [45]. Although activation of cell cycle components by Aβ has been well-documented, some studies have concluded that cell cycle deregulation occurs early in the disease process, prior to the onset of neuropathology [201,382]. Indeed, cell cycle-promoting CDKs phosphorylate APP while their inhibition reduces Aβ production resulting in decreased synapse loss and memory impairment in AD mice [383]. Microglia activated by Aβ can induce cell cycle deregulation and death of co-cultured neurons demonstrating that glial signals contribute to the neurotoxic effects of Aβ by activating CDKs [384]. CDK inhibitors inhibit neuropathological abnormalities and reduce behavioral deficits in AD mice in several studies [40,45,202]. One analysis of the literature reported that 13 of 37 AD-risk genes are likely to be functionally involved in cell cycle or mitosis regulation [36]. In sum and with regards to where in the cell cycle neurons die, there is consensus from cell culture and mouse models (as well as some data from postmortem samples from AD patients) that in the AD brain neurons leave G0 to enter the cell cycle but are unable to traverse through S-phase and complete mitosis resulting in their death by apoptosis [34,35,36,201,385,386]. Some evidence indicates that neurons are arrested at cell cycle checkpoints, most commonly at G2/M, prior to dying by apoptosis, [387,388]. While most evidence points to defects prior to the M-phase in AD models, upregulation of mitotic CDK activators and downregulation of mitotic CDK inhibitors has been observed as well [389,390].
Although aberrant cell cycle entry is the most described effect of activated cell cycle-related CDKs in neurons, other mechanisms by which these CDKs promote neuropathology have been described. For example, activation of CDK1 increases APP processing to Aβ and impairs autophagy in AD through inhibition of Beclin-1 [204]. CDK1 can phosphorylate Aβ at Ser26 increasing its neurotoxicity and reducing its ability to form insoluble fibrils [391,392]. CDK1 and CDK2 also contribute to Tau hyperphosphorylation [201,393,394,395,396]. In cell culture experiments, toxicity by Aβ was shown to involve induction of CDK2 activity and its phosphorylation of Tau [397] (Figure 1C)
The CDK believed to be most involved in AD pathogenesis is CDK5. Proteolysis of p35 (and p39) by an increase in calpain activity produces a fragment, p25 (and p29), that forms a highly stable complex with CDK5 causing its hyperactivation [398]. Because p25 is cytosolic, CDK5 activity is also shifted from the membrane, where p35 is localized, to the cytosol. The mislocalization of CDK5 to the cytoplasm lies results in the phosphorylation of proteins that are not natural substrates of the kinase. As described below, one of these CDK-phosphorylated cytosolic protein is Tau. Indeed, in addition to GSK3, CDK5 is considered to be another major Tau kinase (Figure 1C). Hyperactive CDK5/p25 phosphorylates Tau at sites that are phosphorylated in the brains of AD patients and promotes its dysfunction [209]. Transgenic mice in which p25 is expressed in the forebrain inducible display neurodegeneration, neuroinflammation and cognitive deficits [399,400,401]. Hyperactive CDK5 also increases production of Aβ by both transcriptional mechanisms [205] and by stimulating the activity of BACE1 [206] and PS1 [207] (Figure 1C). An early consequence of the CDK5-mediated increase in Aβ is an impairment of synaptic homeostasis [205]. Increased Aβ can also activate CDK5 and increase Tau hyperphosphorylation suggesting a positive feedback loop [204,402]. Enhanced CDK5 activity in glial cells promotes neuroinflammation by increasing production of lysophosphatidylcholine, which induces release of inflammatory cytokines [212]. Other studies have found that CDK5 stimulated by Aβ exposure activates CDKs1, 2 and 4 by phosphorylating them along with inactivation of phosphatases that negatively regulate these CDKs [377]. Once activated the CDKs phosphorylate lamins (which normally occurs during mitosis in dividing cells) resulting in damage of the nuclear envelope by phosphorylating [208]. It is widely recognized that in AD (and other neurodegenerative diseases) the nuclear membrane is deformed and nucleocytoplasmic transport is dysfunctional [403,404]. It deserves mention that the results of studies by some have disputed the aforementioned p25-mediated CDK5 hyperactivation model initially proposed by Patrick, Cruz and Tsai et al. [369,405]. These laboratories propose that p25 is produced normally, has an important role in memory and learning, and that its expression is actually reduced in the AD brain contributing to cognitive deficits [369,405]. Regardless of whether activation results from p25 production or through other mechanisms, it is widely accepted that CDK5 activity is elevated in the AD brain and is a key contributor of disease pathogenesis [406,407,408]. Interestingly, recent studies have described that like GSK3 and p38 MAPK, CDK5 also inhibits autophagy although acting through a different mechanism involving phosphorylation of the Vps34 protein which interferes with its interaction with Beclin-1, an interaction required for the initiation of autophagy. Besides affecting clearance of pathogenic Aβ and Tau, one study has shown that CDK-mediated inhibition of autophagy deregulates APP processing [204,210,211].

3.4.2. CDKs as a Therapeutic Target in AD

Many studies have described protection by CDK inhibitors against Aβ-induced toxicity in cultured neurons [34,35,36,201,385,386]. CDK inhibitors have also been successfully tested in AD mice. In one mouse study administration of the non-selective CDK inhibitor flavopiridol in mice prevented the increased expression of cell cycle proteins and reduced memory impairment resulting from intrecerebroventricular injection of oligomeric Aβ [45]. Similar findings were described in another study in which administration of another non-selective CDK inhibitor, roscovitine, reversed disease-associated transcriptomic changes, reduced Aβ and Tau pathology, and improved behavioral performance in an AD mouse model [202]. Most emphasis on CDKs as a therapeutic target has been on CDK5. Intracerebroventricular infusions of the CDK5 inhibitory peptide (CIP), a 125 aa peptide generated from p35 by C- and N-terminal cleavage inhibits CDK5/p25 activity and reduces Tau hyperphosphorylation, the number of NFTs and neurodegeneration in the p25-overpressing transgenic mouse model of AD [409]. Other studies in which CIP was delivered through an adeno-associated viral vector was similarly protective in p25-transgenic mice reducing neuroinflammation, neuropathology and neurodegeneration [410,411,412]. Administration of another 20 aa peptide derived from p35 cleavage, TFP5, and also inhibits CDK5/p25 activity, reduces neuropathology and restores synaptic function and cognitive function in both the 5XFAD and p25-overexpressing mice. Although derived from p35 TFP5 does not inhibit endogenous CDK5/p35 activity or the activity of other cell cycle-promoting CDKs [413,414]. Interestingly, and in contrast to CIP, TFP5 was shown to be brain permeable increasing its potential as a therapeutic agent. Pharmacological inhibition of CDK5 prevents the reduction of hippocampal neurogenesis in adult AD mice although the inhibitor used in the study, roscovitine, also inhibits cell cycle-promoting CDKs [415,416]. Another study reported that the anti-diabetes drug pioglitazone, a thiazolidinedione compound, inhibits CDK5 activity by decreasing p35 protein level [417]. Pioglitazone restored LTP in Aβ-treated hippocampal slices and reduced memory deficits in AD mouse models [417]. However, pioglitazone has many effects including regulating insulin signaling and PPARg [418]. Despite these limitations, inhibition of CDK5 activity represents a promising starting point in the development of treatments for AD.

3.4.3. CDKs in HD

Multiple lines of evidence supports a role for deregulated expression of cell cycle components and inappropriate entry of neurons in the cell cycle in the pathogenesis of HD. Expression of mut-Htt is sufficient to cause cell cycle defects in tissue culture models leading to cell death [419,420,421,422]. HD mice and cultured neurons expressing mut-Htt displayed induction of cyclin B activity [422]. In HD mice, reactivation of cell cycle markers was found in early and middle stages of the disease process. Increased expression of E2F1, cyclin D1and cyclin E, all necessary for progression through the G1 phase, has been found in the striatum of HD striatum and HD mice [419,420]. Another study described increased phosphorylation of Rb and reduced expression of the CDK-inhibitory protein, p27, in HD mice and in a cell culture model of HD [423]. Treatment with CDK inhibitors was protective in the cell culture HD model. Using a mouse model and careful examination of the striatum one study reported that cell cycle entry was preceded by perinuclear accumulation of mut-Htt and damage of the nuclear membrane [422]. The induction of cell cycle entry and neuronal death by perinuclear accumulation of mut-Htt was confirmed in cultured cortical neurons [422]. Deregulation of the ER stress response pathway resulting from reduced ATF6α/Rheb (Ras-homologue enriched in brain) signaling is another mechanism that has been proposed to induce cell cycle reactivation and death of neurons [420]. A network analysis of human post-mortem microarrays identified CDK1 as one of 19 genes that were particularly significant to HD pathogenesis [424]. Although most of the cell culture and mouse experiments describe induction of cell cycle proteins operating at G1 and S phases, this finding suggests that deregulation of M-phase CDKs are also important. In sum, although the evidence for cell cycle reactivation in HD is strong, the roles of individual CDKs and cyclins in causing neuronal death is unresolved.
Compelling evidence links deregulated CDK5 activity to HD pathogenesis. CDK5 activity is elevated in HD mice and patients [425] and in mice, contributes to behavioral abnormalities characterizing the disease and neurodegeneration. In HD patients, cognitive disturbances and learning and memory deficits manifest well before motor dysfunction [426,427,428]. In HD mice, the genetic knockdown of CDK5 expression attenuates these progressive cognitive and memory impairments [429]. This recovery was attributed to readjusting levels of specific glutamate receptor subunits and restoring hippocampal spine density [429]. Depressive behavior has been suggested to correlate with cognitive disturbances in HD and reflects severity of impaired cognitive performance [430]. Deregulation of CDK5 activity contributes to depressive-like behavior in HD mice acting through aberrant phosphorylation of DARPP-32 (dopamine- and cAMP-regulated phosphoprotein 32), a protein expressed selectively in the striatum [216]. This suggests that CDK5 mis-regulation could affect multiple HD-associated symptoms. Both genetic knockdown and pharmacological inhibition of CDK5 with roscovitine ameliorates depressive-like behavior in HD mice [216]. A defining feature of HD is involuntary motor movements, which results from degeneration of GABAergic neurons in the striatum. Different laboratories have concluded that elevated CDK5 activity contributes to striatal neurodegeneration. One study described increased production of p25 and elevated CDK5 phosphorylation induces oxidative stress and NMDA receptor activity in striatal neurons increasing their vulnerability to death [214,215]. Oxidative stress and NMDA receptor activation are known to cause excitotoxicity which is considered to be the major mode of neuronal death in HD [65,431]. One study found that CDK5/p25 promotes excitotoxicity by phosphorylation and destabilization of Fbxw7 (F-box/WD repeat-containing protein 7), which results in increased expression of the pro-apoptotic protein, c-jun [217]. Another study described that in response to NMDA receptor stimulation CDK5/p25 phosphorylated and destabilized Cdh1 causing the stabilization and accumulation of cyclin B1 and leading to apoptotic death [218]. Genetic reduction of p25 or p35 in HD mice attenuated CDK5 hyperactivity and protects against NMDA receptor-mediated excitotoxicity [432]. Striatal neurons are sensitive to dopaminergic toxicity [77,433]. CDK5 has been found to increased dopamine neurotoxicity in HD models an action involving increased mitochondrial fission [213]. Pharmacological inhibition of CDK5 reduces mitochondrial fission and protects against dopamine toxicity [213].
It is noteworthy that although the overwhelming consensus is that CDK5 activity is elevated in HD and contributes to disease pathogenesis, two separate studies have described that CDK5 phosphorylates mut-Htt and that this modification protects it from cleavage to the toxic N-terminus fragment [434]. In one of these studies it was found that CDK5 activity was reduced by expression of mut-Htt through direct binding which prevented the interaction of p35 with CDK5. The reduction of CDK5 activity enhanced cleavage of mut-Htt and therefore increased neurotoxicity in neuronal cell lines [434]. Protective phosphorylation of mut-Htt by CDK5 was confirmed in another study in primary striatal neurons from HD mice and subjected to DNA-damage using camptothecin [435]. This study also described that even wild-type Htt (which is also phosphorylated by CDK5) is rendered neurotoxic in response to DNA-damage if CDK5-mediated phosphorylation is inhibited [435]. These authors described a substantial reduction of both CDK5 and p35 levels in the striatum of patients with late-stage HD. The expression of p25 was not studied in the two studies and the issue of whether mut-Htt disrupts CDK binding to p25 is unclear. Therefore, while the overall consensus supports a role for a causal role for elevated CDK5 in HD, more research is necessary to establish this unequivocally.

3.4.4. CDKs Are a Therapeutic Target for HD

Chemical inhibitors of CDK1 and CDK2 protect against neurodegeneration in the 3-NP (3-nitropropionic acid)-induced mouse model of HD [436]. Genetic knockdown of CDK5 protects against corticostriatal learning deficits, hippocampal-dependent memory impairment and depressive behavior in HD mice [216,429]. As indicated above, CDK inhibition provides protection against dopaminergic toxicity and aberrant mitochondrial fission in a cell culture model of HD [213], both that are believed to play key roles in disease pathogenesis [65,437,438]. Results of some studies suggest that whereas CDK5/p25 promotes HD neuropathology, CDK5/p35 may have protective effects. Consistent with this, knockdown or inhibition of CDK5/p35 has been found to have either no effect or a negative effect in HD cells and mice [434,439,440]. Although yet to be rigorously tested and confirmed, this raises the possibility that CDK5-based HD therapeutics will have to selectively target CDK5/p25.

3.5. Would Inhibiting Any One Target Work for AD and HD? The Case for Multi-Target Therapies

Intense effort has been made over the past two decades to develop an effective disease-modifying therapy for AD. Much of this effort has targeted Aβ and Tau. These include active and passive immunotherapy against Aβ and Tau, inhibitors of their oligomerization and fibrillization, inhibitors of Aβ production and enhancers of its clearance, and inhibitors of Tau-phosphorylating kinases and microtubule stabilizers [57]. However, it is now well-accepted that the pathogenic mechanisms underlying AD are complex and involve critical contributions from multiple neuronal and glial cell types and molecular targets within and outside them in addition to Aβ and Tau. Even within vulnerable regions of the AD brain there can be diverse signaling mechanisms that contribute to disease pathogenesis. For example, results from single cell transcriptomic analysis indicate that AD pathology-related gene expression changes can be both cell-specific as well as common across cell types in the brain [441]. The cell-specific changes were found to be diverse. Moreover, the transcriptional profiles were different between sexes in several cell types [441]. A more recent single-nucleus transcriptomic study conducted both tissue from 5XFAD mice and AD patients with TREM2 mutations found that transcriptional signatures in human AD in microglia, astrocytes and oligodendrocytes were strikingly different from those observed in mice [442]. Surprisingly, there was limited concordance between the two aforementioned transcriptomics studies [441,442]. To add to the complexity, recent neuropathological studies indicate multiple subtypes of AD with distinct clinical presentation, age at onset, disease duration, and rate of cognitive decline [443]. A study that utilized PET scanning to study spatiotemporal spreading of Tau in living patients described four distinct spreading and deposition patterns of Tau that presented with distinct cognitive profiles and disease progression patterns [444]. Along with the recognition that the disease process in AD begins many years before Aβ and Tau pathology is detectable suggests that therapies should simultaneously target multiple disease-relevant molecules that act upstream of Aβ and Tau abnormalities.
As described in this review, kinases including GSK3, p38 MAPK and CDKs affect disease initiation and progression in multiple ways acting both upstream and during neuropathology and cognitive impairment. For example, in the case of p38 MAPK, contributions to AD are cell autonomous (stimulating cell death mechanisms within neurons) and non-cell autonomous (through release of cytokines from astrocytes and microglia). A recent study described Tau pathology in oligodendrocytes in AD mice where it co-localized with active p38 MAPK, which regulated Tau seeding [445]. An attractive therapeutic approach would be to identify drugs that target multiple molecules or chemically modify existing single-target drugs to regulate additional targets. Some such drugs, initially believed to be selective again a target, have subsequently been found to inhibit other targets relevant to disease pathology. Examples of such drugs described in this review include lithium, valproic acid, flavopiridol, roscovitine and rapamycin. Kenpaullone, a compound developed and often used as a CDK inhibitor, also inhibits GSK3 and has been utilized in some studies as a GSK3 inhibitor [446,447]. Kenpaullone reduces phosphorylation of APP and lowers its processing to Aβ [447]. Recent efforts in the development of novel treatments for human diseases has led to the synthesis of multi-kinase inhibitors [448]. Rational designing has generated other compounds that inhibit CDK1/GSK3 [449,450], CDK1/CDK5/GSK3 [451,452,453], CDK1/CDK2/CDK5/GSK3 [454] and p38 MAPK/GSK3 [455]. Recently dual CDK5/GSK3 inhibitors possessing a tetrahydropyridine isoindolone skeleton have been identified [456,457]. Using a different approach another CDK5/GSK3 inhibiting compound, LDC8, was identified and shown to protect against neuroinflammation-induced neuronal death in vitro [458]. In comparison, genetic knockdown of CDK5 displayed only partial protection indicating that simultaneous inhibition of both GSK3 and CDK5 is necessary for complete protection. A few of these multi-kinase inhibitors have been tested in model systems as candidates for AD therapeutics. LDC8 was also found to protect against neuroinflammation and synaptic degeneration in a zebrafish model of AD [458]. Screening of benzofuropyridine compounds for their ability to inhibit phosphorylation and oligomerization of Tau has led to the identification of a compound that inhibits GSK3β, CDK1 and CDK5 [396]. The triple-kinase inhibitor was more effective at inhibiting Tau phosphorylation and oligomerization in cultured cells than other compounds identified in the screen that inhibited only one or two of the kinases [396]. Another study reported that dihydroxy-1-aza-9-oxafluorene compounds that inhibit CDK1/GSK-3β/CDK5/p25 robustly inhibit Tau phosphorylation at nanomolar concentrations [452]. Interestingly, a compound, HSB13, which inhibits all the AD-causing kinases covered in this review - GSK3, CDKs1,2 and 5 and p38 MAPK - has been shown to protect against neurodegeneration in a Drosophila model of AD [459]. The potential of multi-kinase inhibitors that target kinases discussed in this review is further elevated by the finding of cross-talk between their signaling actions in the disease process. For example, elevated CDK5 has been proposed to cause neuronal damage and cognitive impairment in mice by stimulating GSK3 signaling [460]. Another study described that GSK3 binds and is activated by p25 [461]. Surprisingly, while binding p25 more effectively than CDK5, GSK3 does not bind to p35. Finally, the action of pharmacological inhibitors, including kinase inhibitors, are easier to titrate than knockdown approaches or immunotherapy. Since most of the molecules that are currently being targeted, including Aβ and Tau and enzymes regulating their production, have important physiological functions, partial inhibition of the targets, which would normalize the pathological increase in activity without completely neutralizing it, could be better than potent inhibition in reducing unwanted actions of treatment. For example, GSK inhibition required for neurological benefit is much lower (20–25% inhibition) than what is required for stabilization of β-catenin, a major cellular target of GSK3 [462,463]. In this regard it is interesting that while inhibiting GSK3, p38 MAPK and CDKS, HSB13, does so partially suggesting that drugs like it might be good candidates for clinical and pre-clinical testing [459].
Although a monogenic disease, HD pathogenesis is also a highly complex process involving many molecules and signaling pathways. Again, targeting multiple disease-relevant molecules simultaneously could represent a more effective approach than the dominant current strategy of reducing mut-Htt levels through antisense technology. Support for this comes from studies in HD mice in which inhibitors of two or more signaling molecules was found to be more effective than inhibition of a single target. For example, in a study using two types of HD mice, benefit on motor and cognitive impairment was observed only when lithium (which inhibits GSK3) and valproic acid (which inhibits HDACs) were co-administered. A study conducted in an HD fly model described that co-treatment with rapamycin (an mTOR inhibitor) and lithium displayed significantly more protection that with either inhibitor alone. The multi-kinase inhibitor, HSB13, displays strong efficacy against neurodegeneration and behavioral deficits in the 3-NP model of HD [459].

4. Future Directions

With the recent advances in in silico drug design (or computer-aided drug design) for neurodegenerative diseases it would be possible to more rapidly design drugs that effectively inhibit two or all three of the key kinases covered in this review [464,465,466]. Development in high-throughput screening platforms that utilize cultured cells or invertebrate models to test large numbers of candidate drugs for multi-kinase inhibition as well as neuroprotection, cytotoxicity and Aβ/Tau/mut-Htt aggregation is also facilitating the development of novel pharmacotherapeutic. Strategies [467,468,469]. Advances in the development of novel preclinical platforms, such as patient-derived iPSC (induced pluripotent stem cells) and 3D organoids, are bridging the translational gap between animal models and human clinical trials [470,471]. An issue that cannot be ignored with regard to therapeutics for neurodegenerative diseases, such as AD and HD, is the delivery of drugs to vulnerable brain parts given the presence of the blood-brain barrier and the large number of efflux transporters in the CNS. Considerable effort has been placed in recent years on the development of nanoparticles as drug delivery vehicles [472,473,474]. A number of nanoparticles, including metal nanoparticles, solid lipid nanoparticles, polymeric nanoparticles, liposomes and extracellular vesicles (EVs) have been used in experimental models of AD and neurodegenerative disease [472,473,474]. Of these polymeric nanoparticles (including hydrogels), liposomes and EVs appear to be particularly attractive because they are more biocompatible and biodegradable. In the case of the three kinases described in this review, nanoparticles can be loaded and encapsulated with chemical inhibitors that reduce enzyme activity, but also biologicals to knockdown their expression including microRNAS (mRNAs) and antisense oligonucleotides [475]. An attractive approach for use in brain pathologies is nasal delivery of lipid nanoparticles and drugs, which has been used in patients for many disorders and ailments, but is now being actively developed for AD and other neurodegenerative diseases [476,477].

Funding

This research received no external funding.

Conflicts of Interest

The author is Founder and CEO of Neugeneron, LLC., a company that seeks to develop therapeutics for neurodegenerative diseases.

References

  1. Holtzman, D.M.; Mandelkow, E.; Selkoe, D.J. Alzheimer disease in 2020. Cold Spring Harb. Perspect. Med. 2012, 2, a011585. [Google Scholar] [CrossRef]
  2. Selkoe, D.J.; Hardy, J. The amyloid hypothesis of Alzheimer’s disease at 25 years. EMBO Mol. Med. 2016, 8, 595–608. [Google Scholar] [CrossRef] [PubMed]
  3. Gallardo, G.; Holtzman, D.M. Amyloid-β and Tau at the Crossroads of Alzheimer’s Disease. Adv. Exp. Med. Biol. 2019, 1184, 187–203. [Google Scholar]
  4. Long, J.M.; Holtzman, D.M. Alzheimer Disease: An Update on Pathobiology and Treatment Strategies. Cell 2019, 179, 312–339. [Google Scholar] [CrossRef]
  5. Zuccato, C.; Cattaneo, E. Huntington’s disease. Handb. Exp. Pharmacol. 2014, 220, 357–409. [Google Scholar]
  6. Walker, F.O. Huntington’s disease. Lancet 2007, 369, 218–228. [Google Scholar] [CrossRef]
  7. Bates, G.P.; Dorsey, R.; Gusella, J.F.; Hayden, M.R.; Kay, C.; Leavitt, B.R.; Nance, M.; Ross, C.A.; Scahill, R.I.; Wetzel, R.; et al. Huntington disease. Nat. Rev. Dis. Primers 2015, 1, 15005. [Google Scholar] [CrossRef]
  8. Cummings, J.; Lee, G.; Ritter, A.; Sabbagh, M.; Zhong, K. Alzheimer’s disease drug development pipeline: 2020. Alzheimers Dement. 2020, 6, e12050. [Google Scholar] [CrossRef]
  9. Revi, M. Alzheimer’s Disease Therapeutic Approaches. Adv. Exp. Med. Biol. 2020, 1195, 105–116. [Google Scholar] [PubMed]
  10. Rodrigues, F.B.; Wild, E.J. Huntington’s Disease Clinical Trials Corner: April 2020. J. Huntingt. Dis. 2020, 9, 185–197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Verheijen, J.; Sleegers, K. Understanding Alzheimer Disease at the Interface between Genetics and Transcriptomics. Trends Genet. 2018, 34, 434–447. [Google Scholar] [CrossRef] [Green Version]
  12. Baranello, R.J.; LBharani, K.; Padmaraju, V.; Chopra, N.; Lahiri, D.K.; HGreig, N.; Pappolla, M.A.; Sambamurti, K. Amyloid-beta protein clearance and degradation (ABCD) pathways and their role in Alzheimer’s disease. Curr. Alzheimer Res. 2015, 12, 32–46. [Google Scholar] [CrossRef] [Green Version]
  13. Nalivaeva, N.N.; Turner, A.J. Targeting amyloid clearance in Alzheimer’s disease as a therapeutic strategy. Br. J. Pharmacol. 2019, 176, 3447–3463. [Google Scholar] [CrossRef]
  14. Sala Frigerio, C.; De Strooper, B. Alzheimer’s Disease Mechanisms and Emerging Roads to Novel Therapeutics. Annu. Rev. Neurosci. 2016, 39, 57–79. [Google Scholar] [CrossRef]
  15. Iqbal, K.; Liu, F.; Gong, C.-X. Tau and neurodegenerative disease: The story so far. Nat. Rev. Neurol. 2016, 12, 15–27. [Google Scholar] [CrossRef]
  16. Gao, Y.; Tan, L.; Yu, J.-T.; Tan, L. Tau in Alzheimer’s Disease: Mechanisms and Therapeutic Strategies. Curr. Alzheimer Res. 2018, 15, 283–300. [Google Scholar] [CrossRef]
  17. Selkoe, D.J. Alzheimer’s disease is a synaptic failure. Science 2002, 298, 789–791. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Scheff, S.W.; Price, D.A. Synaptic pathology in Alzheimer’s disease: A review of ultrastructural studies. Neurobiol. Aging 2003, 24, 1029–1046. [Google Scholar] [CrossRef] [PubMed]
  19. Scheff, S.W.; Price, D.A.; Schmitt, F.A.; Mufson, E.J. Hippocampal synaptic loss in early Alzheimer’s disease and mild cognitive impairment. Neurobiol. Aging 2006, 27, 1372–1384. [Google Scholar] [CrossRef] [PubMed]
  20. Scheff, S.W.; Price, D.A.; Schmitt, F.A.; Roberts, K.N.; Ikonomovic, M.D.; Mufson, E.J. Synapse stability in the precuneus early in the progression of Alzheimer’s disease. J. Alzheimers Dis. 2013, 35, 599–609. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Heneka, M.T.; O’Banion, M.K.; Terwel, D.; Kummer, M.P. Neuroinflammatory processes in Alzheimer’s disease. J. Neural Transm. 2010, 117, 919–947. [Google Scholar] [CrossRef]
  22. Calsolaro, V.; Edison, P. Neuroinflammation in Alzheimer’s disease: Current evidence and future directions. Alzheimers Dement. 2016, 12, 719–732. [Google Scholar] [CrossRef]
  23. Dorszewska, J.; Prendecki, M.; Oczkowska, A.; Dezor, M.; Kozubski, W. Molecular Basis of Familial and Sporadic Alzheimer’s Disease. Curr. Alzheimer Res. 2016, 13, 952–963. [Google Scholar] [CrossRef] [PubMed]
  24. Wolfe, C.M.; Fitz, N.F.; Nam, K.N.; Lefterov, I.; Koldamova, R. The Role of APOE and TREM2 in Alzheimer’s Disease-Current Understanding and Perspectives. Int. J. Mol. Sci. 2018, 20, 81. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Karch, C.M.; Goate, A.M. Alzheimer’s disease risk genes and mechanisms of disease pathogenesis. Biol. Psychiatry 2015, 77, 43–51. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. De Strooper, B.; Karran, E. The Cellular Phase of Alzheimer’s Disease. Cell 2016, 164, 603–615. [Google Scholar] [CrossRef] [Green Version]
  27. Kim, E.; Otgontenger, U.; Jamsranjav, A.; Kim, S.S. Deleterious Alteration of Glia in the Brain of Alzheimer’s Disease. Int. J. Mol. Sci. 2020, 21, 6676. [Google Scholar] [CrossRef] [PubMed]
  28. Carmichael, O.; Schwarz, C.; Drucker, D.; Fletcher, E.; Harvey, D.; Beckett, L.; Jack, C.R.; Weiner, M.; DeCarli, C. Alzheimer’s Disease Neuroimaging Initiative. Longitudinal changes in white matter disease and cognition in the first year of the Alzheimer disease neuroimaging initiative. Arch. Neurol. 2010, 67, 1370–1378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Lu, P.H.; Lee, G.J.; Shapira, J.; Jimenez, E.; Mather, M.J.; Thompson, P.M.; Bartzokis, G.; Mendez, M.F. Regional differences in white matter breakdown between frontotemporal dementia and early-onset Alzheimer’s disease. J. Alzheimers Dis. 2014, 39, 261–269. [Google Scholar] [CrossRef] [Green Version]
  30. Bartzokis, G. Age-related myelin breakdown: A developmental model of cognitive decline and Alzheimer’s disease. Neurobiol. Aging 2004, 25, 5–18. [Google Scholar] [CrossRef]
  31. Bartzokis, G. Alzheimer’s disease as homeostatic responses to age-related myelin breakdown. Neurobiol. Aging 2011, 32, 1341–1371. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Desai, M.K.; Guercio, B.J.; Narrow, W.C.; Bowers, W.J. An Alzheimer’s disease-relevant presenilin-1 mutation augments amyloid-beta-induced oligodendrocyte dysfunction. Glia 2011, 59, 627–640. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Narasimhan, S.; Changolkar, L.; Riddle, D.M.; Kats, A.; Stieber, A.; Weitzman, S.A.; Zhang, B.; Li, Z.; Roberson, E.D.; Trojanowski, J.Q.; et al. Human tau pathology transmits glial tau aggregates in the absence of neuronal tau. J. Exp. Med. 2020, 217. [Google Scholar] [CrossRef] [PubMed]
  34. Frade, J.M.; Ovejero-Benito, M.C. Neuronal cell cycle: The neuron itself and its circumstances. Cell Cycle 2015, 14, 712–720. [Google Scholar] [CrossRef] [Green Version]
  35. Keeney, J.T.; Swomley, A.M.; Harris, J.L.; Fiorini, A.; Mitov, M.I.; Perluigi, M.; Sultana, R.; Butterfield, D.A. Cell cycle proteins in brain in mild cognitive impairment: Insights into progression to Alzheimer disease. Neurotox. Res. 2012, 22, 220–230. [Google Scholar] [CrossRef]
  36. Rao, C.V.; Asch, A.S.; Carr, D.J.J.; Yamada, H.Y. “Amyloid-beta accumulation cycle” as a prevention and/or therapy target for Alzheimer’s disease. Aging Cell 2020, 19, e13109. [Google Scholar] [CrossRef]
  37. Herrup, K. The contributions of unscheduled neuronal cell cycle events to the death of neurons in Alzheimer’s disease. Front. Biosci. 2012, 4, 2101–2109. [Google Scholar] [CrossRef]
  38. Folch, J.; Junyent, F.; Verdaguer, E.; Auladell, C.; Pizarro, J.G.; Beas-Zarate, C.; Pallàs, M.; Camins, A. Role of cell cycle re-entry in neurons: A common apoptotic mechanism of neuronal cell death. Neurotox. Res. 2012, 22, 195–207. [Google Scholar] [CrossRef]
  39. Lee, H.G.; Casadesus, G.; Zhu, X.; Castellani, R.J.; McShea, A.; Perry, G.; Petersen, R.B.; Bajic, V.; Smith, M.A. Cell cycle re-entry mediated neurodegeneration and its treatment role in the pathogenesis of Alzheimer’s disease. Neurochem. Int. 2009, 54, 84–88. [Google Scholar] [CrossRef] [Green Version]
  40. Xiang, Z.; Ho, L.; Valdellon, J.; Borchelt, D.; Kelley, K.; Spielman, L.; Aisen, P.S.; Pasinetti, G.M. Cyclooxygenase (COX)-2 and cell cycle activity in a transgenic mouse model of Alzheimer’s disease neuropathology. Neurobiol. Aging 2002, 23, 327–334. [Google Scholar] [CrossRef]
  41. Lopes, J.P.; Blurton-Jones, M.; Yamasaki, T.R.; Agostinho, P.; LaFerla, F.M. Activation of cell cycle proteins in transgenic mice in response to neuronal loss but not amyloid-beta and tau pathology. J. Alzheimers Dis. 2009, 16, 541–549. [Google Scholar] [CrossRef]
  42. Sultana, R.; Butterfield, D.A. Regional expression of key cell cycle proteins in brain from subjects with amnestic mild cognitive impairment. Neurochem. Res. 2007, 32, 655–662. [Google Scholar] [CrossRef]
  43. Evans, T.A.; Raina, A.K.; Delacourte, A.; Aprelikova, O.; Lee, H.G.; Zhu, X.; Perry, G.; Smith, M.A. BRCA1 may modulate neuronal cell cycle re-entry in Alzheimer disease. Int. J. Med. Sci. 2007, 4, 140–145. [Google Scholar] [CrossRef] [Green Version]
  44. Arendt, T.; Brückner, M.K.; Mosch, B.; Lösche, A. Selective cell death of hyperploid neurons in Alzheimer’s disease. Am. J. Pathol. 2010, 177, 15–20. [Google Scholar] [CrossRef]
  45. Leggio, G.M.; Catania, M.V.; Puzzo, D.; Spatuzza, M.; Pellitteri, R.; Gulisano, W.; Torrisi, S.A.; Giurdanella, G.; Piazza, C.; Impellizzeri, A.R.; et al. The antineoplastic drug flavopiridol reverses memory impairment induced by Amyloid-ß1-42 oligomers in mice. Pharmacol. Res. 2016, 106, 10–20. [Google Scholar] [CrossRef] [PubMed]
  46. 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.; et al. The cholinergic system in the pathophysiology and treatment of Alzheimer’s disease. Brain 2018, 141, 1917–1933. [Google Scholar] [CrossRef] [PubMed]
  47. Uddin, M.S.; Tewari, D.; Sharma, G.; Kabir, M.T.; Barreto, G.E.; Bin-Jumah, M.N.; Perveen, A.; Abdel-Daim, M.M.; Ashraf, G.M. Molecular Mechanisms of ER Stress and UPR in the Pathogenesis of Alzheimer’s Disease. Mol. Neurobiol. 2020, 57, 2902–2919. [Google Scholar] [CrossRef]
  48. Hetz, C.; Saxena, S. ER stress and the unfolded protein response in neurodegeneration. Nat. Rev. Neurol. 2017, 13, 477–491. [Google Scholar] [CrossRef] [PubMed]
  49. Hamano, T.; Hayashi, K.; Shirafuji, N.; Nakamoto, Y. The Implications of Autophagy in Alzheimer’s Disease. Curr. Alzheimer Res. 2018, 15, 1283–1296. [Google Scholar] [CrossRef] [PubMed]
  50. Li, Q.; Liu, Y.; Sun, M. Autophagy and Alzheimer’s Disease. Cell Mol. Neurobiol. 2017, 37, 377–388. [Google Scholar]
  51. Reddy, P.H.; Oliver, D.M. Amyloid Beta and Phosphorylated Tau-Induced Defective Autophagy and Mitophagy in Alzheimer’s Disease. Cells 2019, 8, 488. [Google Scholar] [CrossRef] [Green Version]
  52. Uddin, M.; Stachowiak, A.; Mamun, A.A.; Tzvetkov, N.T.; Takeda, S.; Atanasov, A.G.; Bergantin, L.B.; Abdel-Daim, M.M.; Stankiewicz, A.M. Autophagy and Alzheimer’s Disease: From Molecular Mechanisms to Therapeutic Implications. Front. Aging Neurosci. 2018, 10, 4. [Google Scholar] [CrossRef]
  53. Zare-Shahabadi, A.; Masliah, E.; Johnson, G.V.W.; Rezaei, N. Autophagy in Alzheimer’s disease. Rev. Neurosci. 2015, 26, 385–395. [Google Scholar] [CrossRef] [PubMed]
  54. Swerdlow, R.H. Mitochondria and Mitochondrial Cascades in Alzheimer’s Disease. J. Alzheimers Dis. 2018, 62, 1403–1416. [Google Scholar] [CrossRef] [Green Version]
  55. Tönnies, E.; Trushina, E. Oxidative Stress, Synaptic Dysfunction, and Alzheimer’s Disease. J. Alzheimers Dis. 2017, 57, 1105–1121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. D’Mello, S.R.; Kindy, M.C. Overdosing on iron: Elevated iron and degenerative brain disorders. Exp. Biol. Med. 2020, 245, 1444–1473. [Google Scholar] [CrossRef] [PubMed]
  57. Xie, J.; Liang, R.; Wang, Y.; Huang, J.; Cao, X.; Niu, B. Progress in Target Drug Molecules for Alzheimer’s Disease. Curr. Top. Med. Chem. 2020, 20, 4–36. [Google Scholar] [CrossRef]
  58. Crous-Bou, M.; Minguillón, C.; Gramunt, N.; Molinuevo, J.L. Alzheimer’s disease prevention: From risk factors to early intervention. Alzheimers Res. Ther. 2017, 9, 71. [Google Scholar] [CrossRef]
  59. Fan, D.-Y.; Wang, Y.-J. Early Intervention in Alzheimer’s Disease: How Early is Early Enough? Neurosci. Bull. 2020, 36, 195–197. [Google Scholar] [CrossRef]
  60. Loera Valencia, R.; Cedazo Minguez, A.; Kenigsberg, P.A.; Page, G.; Duarte, A.I.; Giusti, P.; Zusso, M.; Robert, P.; Frisoni, G.B.; Cattaneo, A.; et al. Current and emerging avenues for Alzheimer’s disease drug targets. J. Intern. Med. 2019, 286, 398–437. [Google Scholar] [CrossRef] [Green Version]
  61. Barage, S.H.; Sonawane, K.D. Amyloid cascade hypothesis: Pathogenesis and therapeutic strategies in Alzheimer’s disease. Neuropeptides 2015, 52, 1–18. [Google Scholar] [CrossRef]
  62. Hernández, F.; Avila, J. The role of glycogen synthase kinase 3 in the early stages of Alzheimers’ disease. FEBS Lett. 2008, 582, 3848–3854. [Google Scholar] [CrossRef] [Green Version]
  63. Raina, A.K.; Zhu, X.; Smith, M.A. Alzheimer’s disease and the Cell Cycle. Acta Neurobiol. Exp. 2004, 64, 107–112. [Google Scholar]
  64. Saudou, F.; Humbert, S. The Biology of Huntingtin. Neuron 2016, 89, 910–926. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Gil, J.M.; Rego, A.C. Mechanisms of neurodegeneration in Huntington’s disease. Eur. J. Neurosci. 2008, 27, 2803–2820. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Truant, R.; Atwal, R.S.; Burtnik, A. Nucleocytoplasmic trafficking and transcription effects of huntingtin in Huntington’s disease. Prog. Neurobiol. 2007, 83, 211–227. [Google Scholar] [CrossRef]
  67. Vitet, H.; Brandt, V.; Saudou, F. Traffic signaling: New functions of huntingtin and axonal transport in neurological disease. Curr. Opin. Neurobiol. 2020, 63, 122–130. [Google Scholar] [CrossRef] [PubMed]
  68. Gao, R.; Chakraborty, A.; Geater, C.; Pradhan, S.; Gordon, K.L.; Snowden, J.; Yuan, S.; Dickey, A.S.; Choudhary, S.; Ashizawa, T.; et al. Mutant huntingtin impairs PNKP and ATXN3, disrupting DNA repair and transcription. Elife 2019, 8. [Google Scholar] [CrossRef]
  69. Martin, D.D.O.; Ladha, S.; Ehrnhoefer, D.E.; Hayden, M.R. Autophagy in Huntington disease and huntingtin in Autophagy. Trends Neurosci. 2015, 38, 26–35. [Google Scholar] [CrossRef] [PubMed]
  70. Franco-Iborra, S.; Plaza-Zabala, A.; Montpeyo, M.; Sebastian, D.; Vila, M.; Martinez-Vicente, M. Mutant HTT (huntingtin) impairs mitophagy in a cellular model of Huntington disease. Autophagy 2020, 17, 672–689. [Google Scholar] [CrossRef]
  71. Molina-Calavita, M.; Barnat, M.; Elias, S.; Aparicio, E.; Piel, M.; Humbert, S. Mutant huntingtin affects cortical progenitor cell division and development of the mouse neocortex. J. Neurosci. 2014, 34, 10034–10040. [Google Scholar] [CrossRef]
  72. Rüb, U.; Seidel, K.; Heinsen, H.; Vonsattel, J.P.; den Dunnen, W.F.; Korf, H.W. Huntington’s disease (HD): The neuropathology of a multisystem neurodegenerative disorder of the human Brain. Brain Pathol. 2016, 26, 726–740. [Google Scholar] [CrossRef]
  73. Nucifora, L.G.; Burke, K.A.; Feng, X.; Arbez, N.; Zhu, S.; Miller, J.; Yang, G.; Ratovitski, T.; Delannoy, M.; Muchowski, P.J.; et al. Identification of novel potentially toxic oligomers formed in vitro from mammalian-derived expanded huntingtin exon-1 protein. J. Biol. Chem. 2012, 287, 16017–16028. [Google Scholar] [CrossRef] [Green Version]
  74. Leitman, J.; Ulrich Hartl, F.; Lederkremer, G.Z. Soluble forms of polyQ-expanded huntingtin rather than large aggregates cause endoplasmic reticulum stress. Nat. Commun. 2013, 4, 2753. [Google Scholar] [CrossRef] [Green Version]
  75. Morozova, O.A.; Gupta, S.; Colby, D.W. Prefibrillar huntingtin oligomers isolated from HD brain potently seed amyloid formation. FEBS Lett. 2015, 589, 1897–1903. [Google Scholar] [CrossRef] [Green Version]
  76. Shannon, K.M.; Fraint, A. Therapeutic advances in Huntington’s Disease. Mov. Disord. 2015, 30, 1539–1546. [Google Scholar] [CrossRef] [PubMed]
  77. Sepers, M.D.; Raymond, L.A. Mechanisms of synaptic dysfunction and excitotoxicity in Huntington’s disease. Drug Discov. Today 2014, 19, 990–996. [Google Scholar] [CrossRef] [PubMed]
  78. Carmo, C.; Naia, L.; Lopes, C.; Rego, A.C. Mitochondrial Dysfunction in Huntington’s Disease. Adv. Exp. Med. Biol. 2018, 1049, 59–83. [Google Scholar] [PubMed]
  79. Labbadia, J.; Morimoto, R.I. Huntington’s disease: Underlying molecular mechanisms and emerging concepts. Trends Biochem. Sci. 2013, 38, 378–385. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Hsiao, H.-Y.; Chen, Y.-C.; Chen, H.-M.; Tu, P.-H.; Chern, Y. A critical role of astrocyte-mediated nuclear factor-κB-dependent inflammation in Huntington’s disease. Hum. Mol. Genet. 2013, 22, 1826–1842. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Hsiao, H.-Y.; Chern, Y. Targeting glial cells to elucidate the pathogenesis of Huntington’s disease. Mol. Neurobiol. 2010, 41, 248–255. [Google Scholar] [CrossRef]
  82. Valadão, P.A.C.; Santos, K.B.S.; e Vieira, T.H.F.; e Cordeiro, T.M.; Teixeira, A.L.; Guatimosim, C.; de Miranda, A.S. Inflammation in Huntington’s disease: A few new twists on an old tale. J. Neuroimmunol. 2020, 348, 577380. [Google Scholar] [CrossRef]
  83. Tai, Y.F.; Pavese, N.; Gerhard, A.; Tabrizi, S.J.; Barker, R.A.; Brooks, D.J.; Piccini, P. Imaging microglial activation in Huntington’s disease. Brain Res. Bull. 2007, 72, 148–151. [Google Scholar] [CrossRef]
  84. 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 2007, 130 Pt 7, 1759–1766. [Google Scholar] [CrossRef] [Green Version]
  85. 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] [Green Version]
  86. Thompson, L.M.; Aiken, C.T.; Kaltenbach, L.S.; Agrawal, N.; Illes, K.; Khoshnan, A.; Martinez-Vincente, M.; Arrasate, M.; O’Rourke, J.G.; Khashwji, H.; et al. IKK phosphorylates Huntingtin and targets it for degradation by the proteasome and lysosome. J. Cell Biol. 2009, 187, 1083–1099. [Google Scholar] [CrossRef] [PubMed]
  87. Gu, X.; Greiner, E.R.; Mishra, R.; Kodali, R.; Osmand, A.; Finkbeiner, S.; Steffan, J.S.; Thompson, L.M.; Wetzel, R.; Yang, X.W. Serines 13 and 16 are critical determinants of full-length human mutant huntingtin induced disease pathogenesis in HD mice. Neuron 2009, 64, 828–840. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Aiken, C.T.; Steffan, J.S.; Guerrero, C.M.; Khashwji, H.; Lukacsovich, T.; Simmons, D.; Purcell, J.M.; Menhaji, K.; Zhu, Y.Z.; Green, K.; et al. Phosphorylation of Threonine 3: Implications for huntingtin aggregation and neurotoxicity. J. Biol. Chem. 2009, 284, 29427–29436. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Atwal, R.S.; Desmond, C.R.; Caron, N.; Maiuri, T.; Xia, J.; Sipione, S.; Truant, R. Kinase inhibitors modulate huntingtin cell localization and toxicity. Nat. Chem. Biol. 2011, 7, 453–460. [Google Scholar] [CrossRef] [PubMed]
  90. Bustamante, M.B.; Ansaloni, A.; Pedersen, J.F.; Azzollini, L.; Cariulo, C.; Wang, Z.M.; Petricca, L.; Verani, M.; Puglisi, F.; Park, H.; et al. Detection of huntingtin exon 1 phosphorylation by Phos-Tag SDS-PAGE: Predominant phosphorylation on threonine 3 and regulation by IKKβ. Biochem. Biophys. Res. Commun. 2015, 463, 1317–1322. [Google Scholar] [CrossRef] [PubMed]
  91. Lontay, B.; Kiss, A.; Virág, L.; Tar, K. How Do Post-Translational Modifications Influence the Pathomechanistic Landscape of Huntington’s Disease? A Comprehensive Review. Int. J. Mol. Sci. 2020, 21, 4282. [Google Scholar] [CrossRef]
  92. Sawant, N.; Reddy, P.H. Role of Phosphorylated Tau and Glucose Synthase Kinase 3 Beta in Huntington’s Disease Progression. J. Alzheimer Dis. 2019. [Google Scholar] [CrossRef]
  93. Baskota, S.U.; Lopez, O.L.; Greenamyre, J.T.; Kofler, J. Spectrum of tau pathologies in Huntington’s disease. Lab. Investig. 2019, 99, 1068–1077. [Google Scholar] [CrossRef] [PubMed]
  94. Zerr, I.; Bähr, M. Is there a role of Tau in Huntington’s disease? J. Neurochem. 2016, 139, 9–10. [Google Scholar] [CrossRef] [PubMed]
  95. Cisbani, G.; Maxan, A.; Kordower, J.H.; Planel, E.; Freeman, T.B.; Cicchetti, F. Presence of tau pathology within foetal neural allografts in patients with Huntington’s and Parkinson’s disease. Brain 2017, 140, 2982–2992. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Vuono, R.; Winder-Rhodes, S.; de Silva, R.; Cisbani, G.; Drouin-Ouellet, J. REGISTRY Investigators of the European Huntington’s Disease Network. The role of tau in the pathological process and clinical expression of Huntington’s disease. Brain 2015, 138 Pt 7, 1907–1918. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Fernández-Nogales, M.; Lucas, J.J. Altered Levels and Isoforms of Tau and Nuclear Membrane Invaginations in Huntington’s Disease. Front. Cell Neurosci. 2019, 13, 574. [Google Scholar] [CrossRef]
  98. Fernández-Nogales, M.; Santos-Galindo, M.; Hernández, I.H.; Cabrera, J.R.; Lucas, J.J. Faulty splicing and cytoskeleton abnormalities in Huntington’s disease. Brain Pathol. 2016, 26, 772–778. [Google Scholar] [CrossRef] [PubMed]
  99. Fernández-Nogales, M.; Cabrera, J.R.; Santos-Galindo, M.; Hoozemans, J.J.; Ferrer, I.; Rozemuller, A.J.; Hernández, F.; Avila, J.; Lucas, J.J. Huntington’s disease is a four-repeat tauopathy with tau nuclear rods. Nat. Med. 2014, 20, 881–885. [Google Scholar] [CrossRef]
  100. Maxan, A.; Cicchetti, F. Tau: A Common Denominator and Therapeutic Target for Neurodegenerative Disorders. J. Exp. Neurosci. 2018, 12, 1179069518772380. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Masnata, M.; Salem, S.; de Rus Jacquet, A.; Anwer, M.; Cicchetti, F. Targeting Tau to Treat Clinical Features of Huntington’s Disease. Front. Neurol. 2020, 11, 580732. [Google Scholar] [CrossRef]
  102. Gratuze, M.; Cisbani, G.; Cicchetti, F.; Planel, E. Is Huntington’s disease a tauopathy? Brain 2016, 139 Pt 4, 1014–1025. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Suelves, N.; Kirkham-McCarthy, L.; Lahue, R.S.; Ginés, S. A selective inhibitor of histone deacetylase 3 prevents cognitive deficits and suppresses striatal CAG repeat expansions in Huntington’s disease mice. Sci. Rep. 2017, 7, 6082. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Jia, H.; Wang, Y.; Morris, C.D.; Jacques, V.; Gottesfeld, J.M.; Rusche, J.R.; Thomas, E.A. The Effects of Pharmacological Inhibition of Histone Deacetylase 3 (HDAC3) in Huntington’s Disease Mice. PLoS ONE 2016, 11, e0152498. [Google Scholar] [CrossRef] [PubMed]
  105. Jia, H.; Pallos, J.; Jacques, V.; Lau, A.; Tang, B.; Cooper, A.; Syed, A.; Purcell, J.; Chen, Y.; Sharma, S.; et al. Histone deacetylase (HDAC) inhibitors targeting HDAC3 and HDAC1 ameliorate polyglutamine-elicited phenotypes in model systems of Huntington’s disease. Neurobiol. Dis. 2012, 46, 351–361. [Google Scholar] [CrossRef] [Green Version]
  106. Bardai, F.H.; D’Mello, S.R. Selective toxicity by HDAC3 in neurons: Regulation by Akt and GSK3beta. J. Neurosci. 2011, 31, 1746–1751. [Google Scholar] [CrossRef] [Green Version]
  107. Bardai, F.H.; Verma, P.; Smith, C.; Rawat, V.; Wang, L.; D’Mello, S.R. Disassociation of histone deacetylase-3 from normal huntingtin underlies mutant huntingtin neurotoxicity. J. Neurosci. 2013, 33, 11833–11838. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Tabrizi, S.J.; Leavitt, B.R.; Landwehrmeyer, G.B.; Wild, E.J.; Saft, C.; Barker, R.A.; Blair, N.F.; Craufurd, D.; Priller, J.; Rickards, H.; et al. Targeting Huntingtin Expression in Patients with Huntington’s Disease. N. Engl. J. Med. 2019, 380, 2307–2316. [Google Scholar] [CrossRef] [PubMed]
  109. Estevez-Fraga, C.; Flower, M.D.; Tabrizi, S.J. Therapeutic strategies for Huntington’s disease. Curr. Opin. Neurol. 2020, 33, 508–518. [Google Scholar] [CrossRef] [PubMed]
  110. Nakamori, M.; Mochizuki, H. Targeting Expanded Repeats by Small Molecules in Repeat Expansion Disorders. Mov. Disord. 2021, 36, 298–305. [Google Scholar] [CrossRef]
  111. Nakamori, M.; Panigrahi, G.B.; Lanni, S.; Gall-Duncan, T.; Hayakawa, H.; Tanaka, H.; Luo, J.; Otabe, T.; Li, J.; Sakata, A.; et al. A slipped-CAG DNA-binding small molecule induces trinucleotide-repeat contractions in vivo. Nat. Genet. 2020, 52, 146–159. [Google Scholar] [CrossRef] [PubMed]
  112. Embi, N.; Rylatt, D.B.; Cohen, P. Glycogen synthase kinase-3 from rabbit skeletal muscle. Separation from cyclic-AMP-dependent protein kinase and phosphorylase kinase. Eur. J. Biochem. 1980, 107, 519–527. [Google Scholar] [CrossRef] [PubMed]
  113. Rylatt, D.B.; Embi, N.; Cohen, P. Glycogen synthase kinase-2 from rabbit skeletal muscle is activated by the calcium-dependent regulator protein. FEBS Lett. 1979, 98, 76–80. [Google Scholar] [CrossRef] [Green Version]
  114. Itarte, E.; Huang, K.P. Purification and properties of cyclic AMP-independent glycogen synthase kinase 1 from rabbit skeletal muscle. J. Biol. Chem. 1979, 254, 4052–4057. [Google Scholar] [CrossRef]
  115. Woodgett, J.R. Molecular cloning and expression of glycogen synthase kinase-3/factor A. EMBO J. 1990, 9, 2431–2438. [Google Scholar] [CrossRef] [PubMed]
  116. Force, T.; Woodgett, J.R. Unique and Overlapping Functions of GSK-3 Isoforms in Cell Differentiation and Proliferation and Cardiovascular Development. J. Biol. Chem. 2009, 284, 9643–9647. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Kerkela, R.; Kockeritz, L.; MacAulay, K.; Zhou, J.; Doble, B.W.; Beahm, C.; Greytak, S.; Woulfe, K.; Trivedi, C.M.; Woodgett, J.R.; et al. Deletion of GSK-3β in mice leads to hypertrophic cardiomyopathy secondary to cardiomyoblast hyperproliferation. J. Clin. Investig. 2008, 118, 3609–3618. [Google Scholar] [CrossRef] [Green Version]
  118. Zhou, J.; Freeman, T.A.; Ahmad, F.; Shang, X.; Mangano, E.; Gao, E.; Farber, J.; Wang, Y.; Ma, X.L.; Woodgett, J.; et al. GSK-3α is a central regulator of age-related pathologies in mice. J. Clin. Investig. 2013, 123, 1821–1832. [Google Scholar] [CrossRef]
  119. Draffin, J.E.; Sánchez-Castillo, C.; Fernández-Rodrigo, A.; Sánchez-Sáez, X.; Ávila, J.; Wagner, F.F.; Esteban, J.A. GSK3α, not GSK3β, drives hippocampal NMDAR-dependent LTD via tau-mediated spine anchoring. EMBO J. 2021, 40, e105513. [Google Scholar] [CrossRef] [PubMed]
  120. Shahab, L.; Plattner, F.; Irvine, E.E.; Cummings, D.M.; Edwards, F.A. Dynamic range of GSK3α not GSK3β is essential for bidirectional synaptic plasticity at hippocampal CA3-CA1 synapses. Hippocampus 2014, 24, 1413–1416. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Kaidanovich-Beilin, O.; Lipina, T.V.; Takao, K.; Van Eede, M.; Hattori, S.; Laliberté, C.; Khan, M.; Okamoto, K.; Chambers, J.W.; Fletcher, P.J.; et al. Abnormalities in brain structure and behavior in GSK-3alpha mutant mice. Mol. Brain 2009, 2, 35. [Google Scholar] [CrossRef] [Green Version]
  122. Lee, F.H.F.; Kaidanovich-Beilin, O.; Roder, J.C.; Woodgett, J.R.; Wong, A.H.C. Genetic inactivation of GSK3α rescues spine deficits in Disc1-L100P mutant mice. Schizophr. Res. 2011, 129, 74–79. [Google Scholar] [CrossRef]
  123. Beurel, E.; Grieco, S.F.; Jope, R.S. Glycogen synthase kinase-3 (GSK3): Regulation, actions, and diseases. Pharmacol. Ther. 2015, 148, 114–131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Goñi-Oliver, P.; Avila, J.; Hernández, F. Calpain-mediated truncation of GSK-3 in post-mortem brain samples. J. Neurosci. Res. 2009, 87, 1156–1161. [Google Scholar] [CrossRef] [PubMed]
  125. Goñi-Oliver, P.; Avila, J.; Hernández, F. Calpain regulates N-terminal interaction of GSK-3β with 14-3-3ζ, p53 and PKB but not with axin. Neurochem. Int. 2011, 59, 97–100. [Google Scholar] [CrossRef] [PubMed]
  126. Jin, N.; Yin, X.; Yu, D.; Cao, M.; Gong, C.X.; Iqbal, K.; Ding, F.; Gu, X.; Liu, F. Truncation and activation of GSK-3β by calpain I: A molecular mechanism links to tau hyperphosphorylation in Alzheimer’s disease. Sci. Rep. 2015, 5, 8187. [Google Scholar] [CrossRef] [Green Version]
  127. Goñi-Oliver, P.; Lucas, J.J.; Avila, J.; Hernández, F. N-terminal cleavage of GSK-3 by calpain: A new form of GSK-3 regulation. J. Biol. Chem. 2007, 282, 22406–22413. [Google Scholar] [CrossRef] [Green Version]
  128. Kandasamy, A.D.; Schulz, R. Glycogen synthase kinase-3beta is activated by matrix metalloproteinase-2 mediated proteolysis in cardiomyoblasts. Cardiovasc. Res. 2009, 83, 698–706. [Google Scholar] [CrossRef] [Green Version]
  129. Feijs, K.L.; Kleine, H.; Braczynski, A.; Forst, A.H.; Herzog, N.; Verheugd, P.; Linzen, U.; Kremmer, E.; Lüscher, B. ARTD10 substrate identification on protein microarrays: Regulation of GSK3β by mono-ADP-ribosylation. Cell Commun. Signal. 2013, 11, 5. [Google Scholar] [CrossRef] [Green Version]
  130. Rosenthal, F.; Feijs, K.L.; Frugier, E.; Bonalli, M.; Forst, A.H.; Imhof, R.; Winkler, H.C.; Fischer, D.; Caflisch, A.; Hassa, P.O.; et al. Macrodomain-containing proteins are new mono-ADP-ribosylhydrolases. Nat. Struct. Mol. Biol. 2013, 20, 502–507. [Google Scholar] [CrossRef]
  131. Sarikhani, M.; Mishra, S.; Maity, S.; Kotyada, C.; Wolfgeher, D.; Gupta, M.P.; Singh, M.; Sundaresan, N.R. SIRT2 deacetylase regulates the activity of GSK3 isoforms independent of inhibitory phosphorylation. Elife 2018, 7, e32952. [Google Scholar] [CrossRef]
  132. Monteserin-Garcia, J.; Al-Massadi, O.; Seoane, L.M.; Alvarez, C.V.; Shan, B.; Stalla, J.; Paez-Pereda, M.; Casanueva, F.F.; Stalla, G.K.; Theodoropoulou, M. Sirt1 inhibits the transcription factor CREB to regulate pituitary growth hormone synthesis. FASEB J. 2013, 27, 1561–1571. [Google Scholar] [CrossRef] [PubMed]
  133. Stadler, S.C.; Vincent, C.T.; Fedorov, V.D.; Patsialou, A.; Cherrington, B.D.; Wakshlag, J.J.; Mohanan, S.; Zee, B.M.; Zhang, X.; Garcia, B.A.; et al. Dysregulation of PAD4-mediated citrullination of nuclear GSK3β activates TGF-β signaling and induces epithelial-to-mesenchymal transition in breast cancer cells. Proc. Natl. Acad. Sci. USA 2013, 110, 11851–11856. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Schenkel, J. Activation of the c-Jun transcription factor following neurodegeneration in vivo. Neurosci. Lett. 2004, 361, 36–39. [Google Scholar] [CrossRef]
  135. Ham, J.; Eilers, A.; Whitfield, J.; Neame, S.J.; Shah, B. c-Jun and the transcriptional control of neuronal apoptosis. Biochem. Pharmacol. 2000, 60, 1015–1021. [Google Scholar] [CrossRef]
  136. D’Mello, S.R. Histone deacetylase-3: Friend and foe of the brain. Exp. Biol. Med. 2020, 245, 1130–1141. [Google Scholar] [CrossRef] [PubMed]
  137. Mielcarek, M.; Zielonka, D.; Carnemolla, A.; Marcinkowski, J.T.; Guidez, F. HDAC4 as a potential therapeutic target in neurodegenerative diseases: A summary of recent achievements. Front. Cell Neurosci. 2015, 9, 42. [Google Scholar] [CrossRef] [Green Version]
  138. Jimenez-Pacheco, A.; Franco, J.M.; Lopez, S.; Gomez-Zumaquero, J.M.; Magdalena Leal-Lasarte, M.; Caballero-Hernandez, D.E.; Cejudo-Guillén, M.; Pozo, D. Epigenetic Mechanisms of Gene Regulation in Amyotrophic Lateral Sclerosis. Adv. Exp. Med. Biol. 2017, 978, 255–275. [Google Scholar]
  139. Linding, R.; Jensen, L.J.; Ostheimer, G.J.; van Vugt, M.A.; Jørgensen, C.; Miron, I.M.; Diella, F.; Colwill, K.; Taylor, L.; Elder, K.; et al. Systematic discovery of in vivo phosphorylation networks. Cell 2007, 129, 1415–1426. [Google Scholar] [CrossRef] [Green Version]
  140. Sutherland, C. What Are the bona fide GSK3 Substrates? Int. J. Alzheimers Dis. 2011, 2011, 505607. [Google Scholar]
  141. Hur, E.-M.; Zhou, F.-Q. GSK3 signalling in neural development. Nat. Rev. Neurosci. 2010, 11, 539–551. [Google Scholar] [CrossRef] [Green Version]
  142. Morales-Garcia, J.A.; Luna-Medina, R.; Alonso-Gil, S.; Sanz-SanCristobal, M.; Palomo, V.; Gil, C.; Santos, A.; Martinez, A.; Perez-Castillo, A. Glycogen synthase kinase 3 inhibition promotes adult hippocampal neurogenesis in vitro and in vivo. ACS Chem. Neurosci. 2012, 3, 963–971. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Moreno-Jiménez, E.P.; Flor-García, M.; Terreros-Roncal, J.; Rábano, A.; Cafini, F.; Pallas-Bazarra, N.; Ávila, J.; Llorens-Martín, M. Adult hippocampal neurogenesis is abundant in neurologically healthy subjects and drops sharply in patients with Alzheimer’s disease. Nat. Med. 2019, 25, 554–560. [Google Scholar] [CrossRef]
  144. Sirerol Piquer, M.; Gomez Ramos, P.; Hernández, F.; Perez, M.; Morán, M.A.; Fuster Matanzo, A.; Lucas, J.J.; Avila, J.; García Verdugo, J.M. GSK3β overexpression induces neuronal death and a depletion of the neurogenic niches in the dentate gyrus. Hippocampus 2011, 21, 910–922. [Google Scholar]
  145. Hernández, F.; Borrell, J.; Guaza, C.; Avila, J.; Lucas, J.J. Spatial learning deficit in transgenic mice that conditionally over-express GSK-3beta in the brain but do not form tau filaments. J. Neurochem. 2002, 83, 1529–1533. [Google Scholar] [CrossRef] [Green Version]
  146. Peineau, S.; Taghibiglou, C.; Bradley, C.; Wong, T.P.; Liu, L.; Lu, J.; Lo, E.; Wu, D.; Saule, E.; Bouschet, T.; et al. LTP inhibits LTD in the hippocampus via regulation of GSK3beta. Neuron 2007, 53, 703–717. [Google Scholar] [CrossRef] [Green Version]
  147. Peineau, S.; Bradley, C.; Taghibiglou, C.; Doherty, A.; Bortolotto, Z.A.; Wang, Y.T.; Collingridge, G.L. The role of GSK-3 in synaptic plasticity. Br. J. Pharmacol. 2008, 153 (Suppl. 1), S428–S437. [Google Scholar] [CrossRef] [Green Version]
  148. Bradley, C.A.; Peineau, S.; Taghibiglou, C.; Nicolas, C.S.; Whitcomb, D.J.; Bortolotto, Z.A.; Kaang, B.K.; Cho, K.; Wang, Y.T.; Collingridge, G.L. A pivotal role of GSK-3 in synaptic plasticity. Front. Mol. Neurosci. 2012, 5, 13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Yuskaitis, C.J.; Jope, R.S. Glycogen synthase kinase-3 regulates microglial migration, inflammation, and inflammation-induced neurotoxicity. Cell Signal. 2009, 21, 264–273. [Google Scholar] [CrossRef] [Green Version]
  150. Jope, R.S.; Yuskaitis, C.J.; Beurel, E. Glycogen synthase kinase-3 (GSK3): Inflammation, diseases, and therapeutics. Neurochem. Res. 2007, 32, 577–595. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Koistinaho, J.; Malm, T.; Goldsteins, G. Glycogen synthase kinase-3β: A mediator of inflammation in Alzheimer’s disease? Int. J. Alzheimers Dis. 2011, 2011, 129753. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Chin, P.C.; Majdzadeh, N.; D’Mello, S.R. Inhibition of GSK3beta is a common event in neuroprotection by different survival factors. Brain Res. Mol. Brain Res. 2005, 137, 193–201. [Google Scholar] [CrossRef] [PubMed]
  153. Thotala, D.K.; Hallahan, D.E.; Yazlovitskaya, E.M. Glycogen synthase kinase 3β inhibitors protect hippocampal neurons from radiation-induced apoptosis by regulating MDM2-p53 pathway. Cell Death Differ. 2012, 19, 387–396. [Google Scholar] [CrossRef] [PubMed]
  154. Takadera, T.; Ohyashiki, T. Glycogen synthase kinase-3 inhibitors prevent caspase-dependent apoptosis induced by ethanol in cultured rat cortical neurons. Eur. J. Pharmacol. 2004, 499, 239–245. [Google Scholar] [CrossRef]
  155. Lucas, J.J.; Hernández, F.; Gómez-Ramos, P.; Morán, M.A.; Hen, R.; Avila, J. Decreased nuclear beta-catenin, tau hyperphosphorylation and neurodegeneration in GSK-3beta conditional transgenic mice. EMBO J. 2001, 20, 27–39. [Google Scholar] [CrossRef]
  156. Leroy, K.; Boutajangout, A.; Authelet, M.; Woodgett, J.R.; Anderton, B.H.; Brion, J.-P. The active form of glycogen synthase kinase-3beta is associated with granulovacuolar degeneration in neurons in Alzheimer’s disease. Acta Neuropathol. 2002, 103, 91–99. [Google Scholar]
  157. Ferrer, I.; Barrachina, M.; Puig, B. Glycogen synthase kinase-3 is associated with neuronal and glial hyperphosphorylated tau deposits in Alzheimer’s disease, Pick’s disease, progressive supranuclear palsy and corticobasal degeneration. Acta Neuropathol. 2002, 104, 583–591. [Google Scholar] [CrossRef] [PubMed]
  158. Himmelstein, D.S.; Ward, S.M.; Lancia, J.K.; Patterson, K.R.; Binder, L.I. Tau as a therapeutic target in neurodegenerative disease. Pharmacol. Ther. 2012, 136, 8–22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Tomizawa, K.; Omori, A.; Ohtake, A.; Sato, K.; Takahashi, M. Tau-tubulin kinase phosphorylates tau at Ser-208 and Ser-210, sites found in paired helical filament-tau. FEBS Lett. 2001, 492, 221–227. [Google Scholar] [CrossRef] [Green Version]
  160. Polydoro, M.; Acker, C.M.; Duff, K.; Castillo, P.E.; Davies, P. Age-dependent impairment of cognitive and synaptic function in the htau mouse model of tau pathology. J. Neurosci. 2009, 29, 10741–10749. [Google Scholar] [CrossRef] [Green Version]
  161. Wen, Y.; Planel, E.; Herman, M.; Figueroa, H.Y.; Wang, L.; Liu, L.; Lau, L.F.; Yu, W.H.; Duff, K.E. Interplay between cyclin-dependent kinase 5 and glycogen synthase kinase 3 beta mediated by neuregulin signaling leads to differential effects on tau phosphorylation and amyloid precursor protein processing. J. Neurosci. 2008, 28, 2624–2632. [Google Scholar] [CrossRef]
  162. Phiel, C.J.; Wilson, C.A.; Lee, V.M.-Y.; Klein, P.S. GSK-3alpha regulates production of Alzheimer’s disease amyloid-beta peptides. Nature 2003, 423, 435–439. [Google Scholar] [CrossRef] [PubMed]
  163. Ryder, J.; Su, Y.; Liu, F.; Li, B.; Zhou, Y.; Ni, B. Divergent roles of GSK3 and CDK5 in APP processing. Biochem. Biophys. Res. Commun. 2003, 312, 922–929. [Google Scholar] [CrossRef]
  164. Su, Y.; Ryder, J.; Li, B.; Wu, X.; Fox, N.; Solenberg, P.; Brune, K.; Paul, S.; Zhou, Y.; Liu, F.; et al. Lithium, a common drug for bipolar disorder treatment, regulates amyloid-beta precursor protein processing. Biochemistry 2004, 43, 6899–6908. [Google Scholar] [CrossRef]
  165. Uemura, K.; Lill, C.M.; Banks, M.; Asada, M.; Aoyagi, N.; Ando, K.; Kubota, M.; Kihara, T.; Nishimoto, T.; Sugimoto, H.; et al. N-cadherin-based adhesion enhances Abeta release and decreases Abeta42/40 ratio. J. Neurochem. 2009, 108, 350–360. [Google Scholar] [CrossRef] [Green Version]
  166. Maesako, M.; Uemura, K.; Kuzuya, A.; Sasaki, K.; Asada, M.; Watanabe, K.; Ando, K.; Kubota, M.; Akiyama, H.; Takahashi, R.; et al. Gain of function by phosphorylation in Presenilin 1-mediated regulation of insulin signaling. J. Neurochem. 2012, 121, 964–973. [Google Scholar] [CrossRef]
  167. Maesako, M.; Uemura, K.; Kubota, M.; Hiyoshi, K.; Ando, K.; Kuzuya, A.; Kihara, T.; Asada, M.; Akiyama, H.; Kinoshita, A. Effect of glycogen synthase kinase 3 β-mediated presenilin 1 phosphorylation on amyloid β production is negatively regulated by insulin receptor cleavage. Neuroscience 2011, 177, 298–307. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Chen, C.H.; Zhou, W.; Liu, S.; Deng, Y.; Cai, F.; Tone, M.; Tone, Y.; Tong, Y.; Song, W. Increased NF-κB signalling up-regulates BACE1 expression and its therapeutic potential in Alzheimer’s disease. Int. J. Neuropsychopharmacol. 2012, 15, 77–90. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Ly, P.T.; Wu, Y.; Zou, H.; Wang, R.; Zhou, W.; Kinoshita, A.; Zhang, M.; Yang, Y.; Cai, F.; Woodgett, J.; et al. Inhibition of GSK3β-mediated BACE1 expression reduces Alzheimer-associated phenotypes. J. Clin. Investig. 2013, 123, 224–235. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Takashima, A.; Honda, T.; Yasutake, K.; Michel, G.; Murayama, O.; Murayama, M.; Ishiguro, K.; Yamaguchi, H. Activation of tau protein kinase I/glycogen synthase kinase-3beta by amyloid beta peptide (25–35) enhances phosphorylation of tau in hippocampal neurons. Neurosci. Res. 1998, 31, 317–323. [Google Scholar] [CrossRef]
  171. Ferrari, A.; Hoerndli, F.; Baechi, T.; Nitsch, R.M.; Götz, J. β-Amyloid induces paired helical filament-like tau filaments in tissue culture. J. Biol. Chem. 2003, 278, 40162–40168. [Google Scholar] [CrossRef] [Green Version]
  172. Yang, C.; Li, X.; Zhang, L.; Li, Y.; Li, L.; Zhang, L. Cornel iridoid glycoside induces autophagy to protect against tau oligomer neurotoxicity induced by the activation of glycogen synthase kinase-3β. J. Nat. Med. 2019, 73, 717–726. [Google Scholar] [CrossRef] [PubMed]
  173. Parr, C.; Carzaniga, R.; Gentleman, S.M.; Van Leuven, F.; Walter, J.; Sastre, M. Glycogen synthase kinase 3 inhibition promotes lysosomal biogenesis and autophagic degradation of the amyloid-β precursor protein. Mol. Cell Biol. 2012, 32, 4410–4418. [Google Scholar] [CrossRef] [Green Version]
  174. Jope, R.S.; Cheng, Y.; Lowell, J.A.; Worthen, R.J.; Sitbon, Y.H.; Beurel, E. Stressed and Inflamed, Can GSK3 Be Blamed? Trends Biochem. Sci. 2017, 42, 180–192. [Google Scholar] [CrossRef] [Green Version]
  175. Hooper, C.; Markevich, V.; Plattner, F.; Killick, R.; Schofield, E.; Engel, T.; Hernandez, F.; Anderton, B.; Rosenblum, K.; Bliss, T.; et al. Glycogen synthase kinase-3 inhibition is integral to long-term potentiation. Eur. J. Neurosci. 2007, 25, 81–86. [Google Scholar] [CrossRef]
  176. Sereno, L.; Coma, M.; Rodriguez, M.; Sanchez-Ferrer, P.; Sanchez, M.B.; Gich, I.; Agullo, J.M.; Perez, M.; Avila, J.; Guardia-Laguarta, C.; et al. A novel GSK-3beta inhibitor reduces Alzheimer’s pathology and rescues neuronal loss in vivo. Neurobiol. Dis. 2009, 35, 359–367. [Google Scholar] [CrossRef] [PubMed]
  177. Farr, S.A.; Ripley, J.L.; Sultana, R.; Zhang, Z.; Niehoff, M.L.; Platt, T.L.; Murphy, M.P.; Morley, J.E.; Kumar, V.; Butterfield, D.A. Antisense oligonucleotide against GSK-3β in brain of SAMP8 mice improves learning and memory and decreases oxidative stress: Involvement of transcription factor Nrf2 and implications for Alzheimer disease. Free Radic. Biol. Med. 2014, 67, 387–395. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Real, S.; Espada, L.; Espinet, C.; Santidrián, A.F.; Tauler, A. Study of the in vivo phosphorylation of E2F1 on Ser403. Biochim. Biophys. Acta 2010, 1803, 912–918. [Google Scholar] [CrossRef] [Green Version]
  179. Espada, L.; Udapudi, B.; Podlesniy, P.; Fabregat, I.; Espinet, C.; Tauler, A. Apoptotic action of E2F1 requires glycogen synthase kinase 3-beta activity in PC12 Cells. J. Neurochem. 2007, 102, 2020–2028. [Google Scholar] [CrossRef] [PubMed]
  180. L’episcopo, F.; Drouin-Ouellet, J.; Tirolo, C.; Pulvirenti, A.; Giugno, R.; Testa, N.; Caniglia, S.; Serapide, M.F.; Cisbani, G.; Barker, R.A.; et al. GSK-3β-induced Tau pathology drives hippocampal neuronal cell death in Huntington’s disease: Involvement of astrocyte-neuron interactions. Cell Death Dis. 2016, 7, e2206. [Google Scholar] [CrossRef] [Green Version]
  181. Falcicchia, C.; Tozzi, F.; Arancio, O.; Watterson, D.M.; Origlia, N. Involvement of p38 MAPK in Synaptic Function and Dysfunction. Int. J. Mol. Sci. 2020, 21, 5624. [Google Scholar] [CrossRef]
  182. Kheiri, G.; Dolatshahi, M.; Rahmani, F.; Rezaei, N. Role of p38/MAPKs in Alzheimer’s disease: Implications for amyloid beta toxicity targeted therapy. Rev. Neurosci. 2018, 30, 9–30. [Google Scholar] [CrossRef]
  183. Kim, S.H.; Smith, C.J.; Van Eldik, L.J. Importance of MAPK pathways for microglial pro-inflammatory cytokine IL-1 beta production. Neurobiol. Aging 2004, 25, 431–439. [Google Scholar] [CrossRef]
  184. de la Torre, A.V.; Junyent, F.; Folch, J.; Pelegrí, C.; Vilaplana, J.; Auladell, C.; Beas-Zarate, C.; Pallàs, M.; Verdaguer, E.; Camins, A. PI3 k/akt inhibition induces apoptosis through p38 activation in neurons. Pharmacol. Res. 2013, 70, 116–125. [Google Scholar] [CrossRef]
  185. Guo, X.-Q.; Cao, Y.-L.; Zhao, L.; Zhang, X.; Yan, Z.-R.; Chen, W.-M. p38 mitogen-activated protein kinase gene silencing rescues rat hippocampal neurons from ketamine-induced apoptosis: An in vitro study. Int. J. Mol. Med. 2018, 42, 1401–1410. [Google Scholar]
  186. Ramiro-Cortés, Y.; Guemez-Gamboa, A.; Morán, J. Reactive oxygen species participate in the p38-mediated apoptosis induced by potassium deprivation and staurosporine in cerebellar granule neurons. Int. J. Biochem. Cell Biol. 2011, 43, 1373–1382. [Google Scholar] [CrossRef]
  187. Feijoo, C.; Campbell, D.G.; Jakes, R.; Goedert, M.; Cuenda, A. Evidence that phosphorylation of the microtubule-associated protein Tau by SAPK4/p38delta at Thr50 promotes microtubule assembly. J. Cell Sci. 2005, 118 Pt 2, 397–408. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Goedert, M.; Hasegawa, M.; Jakes, R.; Lawler, S.; Cuenda, A.; Cohen, P. Phosphorylation of microtubule-associated protein tau by stress-activated protein kinases. FEBS Lett. 1997, 409, 57–62. [Google Scholar] [CrossRef]
  189. Maphis, N.; Jiang, S.; Xu, G.; Kokiko-Cochran, O.N.; Roy, S.M.; Van Eldik, L.J.; Watterson, D.M.; Lamb, B.T.; Bhaskar, K. Selective suppression of the α isoform of p38 MAPK rescues late-stage tau pathology. Alzheimer Res. Ther. 2016, 8, 54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Delston, R.B.; Matatall, K.A.; Sun, Y.; Onken, M.D.; Harbour, J.W. p38 phosphorylates Rb on Ser567 by a novel, cell cycle-independent mechanism that triggers Rb-Hdm2 interaction and apoptosis. Oncogene 2011, 30, 588–599. [Google Scholar] [CrossRef] [Green Version]
  191. Wang, S.; Nath, N.; Minden, A.; Chellappan, S. Regulation of Rb and E2F by signal transduction cascades: Divergent effects of JNK1 and p38 kinases. EMBO J. 1999, 18, 1559–1570. [Google Scholar] [CrossRef] [Green Version]
  192. Rutigliano, G.; Stazi, M.; Arancio, O.; Watterson, D.M.; Origlia, N. An isoform-selective p38α mitogen-activated protein kinase inhibitor rescues early entorhinal cortex dysfunctions in a mouse model of Alzheimer’s disease. Neurobiol. Aging 2018, 70, 86–91. [Google Scholar] [CrossRef]
  193. Origlia, N.; Righi, M.; Capsoni, S.; Cattaneo, A.; Fang, F.; Stern, D.M.; Chen, J.X.; Schmidt, A.M.; Arancio, O.; Du Yan, S.; et al. Receptor for Advanced Glycation End Product-Dependent Activation of p38 Mitogen-Activated Protein Kinase Contributes to Amyloid-β-Mediated Cortical Synaptic Dysfunction. J. Neurosci. 2008, 28, 3521–3530. [Google Scholar] [CrossRef] [Green Version]
  194. Wang, Q.; Walsh, D.M.; Rowan, M.J.; Selkoe, D.J.; Anwyl, R. Block of Long-Term Potentiation by Naturally Secreted and Synthetic Amyloid β-Peptide in Hippocampal Slices Is Mediated via Activation of the Kinases c-Jun N-Terminal Kinase, Cyclin-Dependent Kinase 5, and p38 Mitogen-Activated Protein Kinase as well as Metabotropic Glutamate Receptor Type 5. J. Neurosci. 2004, 24, 3370–3378. [Google Scholar]
  195. Culbert, A.A.; Skaper, S.D.; Howlett, D.R.; Evans, N.A.; Facci, L.; Soden, P.E.; Seymour, Z.M.; Guillot, F.; Gaestel, M.; Richardson, J.C. P38 MAPK is involved in enhanced NMDA receptor-dependent excitotoxicity in YAC transgenic mouse model of Huntington disease. Neurobiol. Dis. 2012, 45, 999–1009. [Google Scholar]
  196. Yang, X.; Zhang, H.; Wu, J.; Yin, L.; Yan, L.-J.; Zhang, C. Humanin Attenuates NMDA-Induced Excitotoxicity by Inhibiting ROS-dependent JNK/p38 MAPK Pathway. Int. J. Mol. Sci. 2018, 19, 2982. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Rivera-Cervantes, M.C.; Castañeda-Arellano, R.; Castro-Torres, R.D.; Gudiño-Cabrera, G.; y Velasco, A.I.F.; Camins, A.; Beas-Zárate, C. P38 MAPK inhibition protects against glutamate neurotoxicity and modifies NMDA and AMPA receptor subunit expression. J. Mol. Neurosci. 2015, 55, 596–608. [Google Scholar] [CrossRef]
  198. McDonald, D.R.; Bamberger, M.E.; Combs, C.K.; Landreth, G.E. β-Amyloid fibrils activate parallel mitogen-activated protein kinase pathways in microglia and THP1 monocytes. J. Neurosci. 1998, 18, 4451–4460. [Google Scholar] [CrossRef] [PubMed]
  199. He, Y.; She, H.; Zhang, T.; Xu, H.; Cheng, L.; Yepes, M.; Zhao, Y.; Mao, Z. p38 MAPK inhibits autophagy and promotes microglial inflammatory responses by phosphorylating ULK1. J. Cell Biol. 2017, 217, 315–328. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  200. Yao, L.; Zhu, Z.; Wu, J.; Zhang, Y.; Zhang, H.; Sun, X.; Qian, C.; Wang, B.; Xie, L.; Zhang, S.; et al. MicroRNA-124 regulates the expression of p62/p38 and promotes autophagy in the inflammatory pathogenesis of Parkinson’s disease. FASEB J. 2019, 33, 8648–8665. [Google Scholar] [CrossRef]
  201. Neve, R.L.; McPhie, D.L. The cell cycle as a therapeutic target for Alzheimer’s disease. Pharmacol. Ther. 2006, 111, 99–113. [Google Scholar] [CrossRef]
  202. Huang, F.; Wang, M.; Liu, R.; Wang, J.Z.; Schadt, E.; Haroutunian, V.; Katsel, P.; Zhang, B.; Wang, X. CDT2-controlled cell cycle reentry regulates the pathogenesis of Alzheimer’s disease. Alzheimers Dement. 2019, 15, 217–231. [Google Scholar] [CrossRef]
  203. Hradek, A.C.; Lee, H.P.; Siedlak, S.L.; Torres, S.L.; Jung, W.; Han, A.H.; Lee, H.G. Distinct chronology of neuronal cell cycle re-entry and tau pathology in the 3xTg-AD mouse model and Alzheimer’s disease patients. J. Alzheimers Dis. 2015, 43, 57–65. [Google Scholar] [CrossRef] [Green Version]
  204. Salminen, A.; Kaarniranta, K.; Kauppinen, A.; Ojala, J.; Haapasalo, A.; Soininen, H.; Hiltunen, M. Impaired autophagy and APP processing in Alzheimer’s disease: The potential role of Beclin 1 interactome. Prog. Neurobiol. 2013, 106–107, 33–54. [Google Scholar] [CrossRef]
  205. Sheng, Y.; Zhang, L.; Su, S.C.; Tsai, L.-H.; Julius Zhu, J. Cdk5 is a New Rapid Synaptic Homeostasis Regulator Capable of Initiating the Early Alzheimer-Like Pathology. Cereb. Cortex. 2016, 26, 2937–2951. [Google Scholar] [CrossRef] [Green Version]
  206. Song, W.-J.; Son, M.-Y.; Lee, H.-W.; Seo, H.; Kim, J.H.; Chung, S.-H. Enhancement of BACE1 Activity by p25/Cdk5-Mediated Phosphorylation in Alzheimer’s Disease. PLoS ONE 2015, 10, e0136950. [Google Scholar] [CrossRef] [PubMed]
  207. Lau, K.F.; Howlett, D.R.; Kesavapany, S.; Standen, C.L.; Dingwall, C.; McLoughlin, D.M.; Miller, C.C. Cyclin-dependent kinase-5/p35 phosphorylates Presenilin 1 to regulate carboxy-terminal fragment stability. Mol. Cell Neurosci. 2002, 20, 13–20. [Google Scholar] [CrossRef]
  208. Chang, K.H.; Multani, P.S.; Sun, K.H.; Vincent, F.; de Pablo, Y.; Ghosh, S.; Gupta, R.; Lee, H.P.; Lee, H.G.; Smith, M.A.; et al. Nuclear envelope dispersion triggered by deregulated Cdk5 precedes neuronal death. Mol. Biol. Cell 2011, 22, 1452–1462. [Google Scholar] [CrossRef]
  209. Kimura, T.; Ishiguro, K.; Hisanaga, S.-I. Physiological and pathological phosphorylation of tau by Cdk5. Front. Mol. Neurosci. 2014, 7, 65. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Furuya, T.; Kim, M.; Lipinski, M.; Li, J.; Kim, D.; Lu, T.; Shen, Y.; Rameh, L.; Yankner, B.; Tsai, L.H.; et al. Negative regulation of Vps34 by Cdk mediated phosphorylation. Mol. Cell 2010, 38, 500–511. [Google Scholar] [CrossRef] [Green Version]
  211. Gupta, K.K.; Singh, S.K. Cdk5: A main culprit in neurodegeneration. Int. J. Neurosci. 2019, 129, 1192–1197. [Google Scholar]
  212. Sundaram, J.R.; Chan, E.S.; Poore, C.P.; Pareek, T.K.; Cheong, W.F.; Shui, G.; Tang, N.; Low, C.M.; Wenk, M.R.; Kesavapany, S. Cdk5/p25-induced cytosolic PLA2-mediated lysophosphatidylcholine production regulates neuroinflammation and triggers neurodegeneration. J. Neurosci. 2012, 32, 1020–1034. [Google Scholar] [CrossRef]
  213. Cherubini, M.; Puigdellívol, M.; Alberch, J.; Ginés, S. Cdk5-mediated mitochondrial fission: A key player in dopaminergic toxicity in Huntington’s disease. Biochim. Biophys. Acta 2015, 1852 Pt A, 2145–2160. [Google Scholar] [CrossRef] [Green Version]
  214. Tang, T.-S.; Chen, X.; Liu, J.; Bezprozvanny, I. Dopaminergic signaling and striatal neurodegeneration in Huntington’s disease. J. Neurosci. 2007, 27, 7899–7910. [Google Scholar] [CrossRef] [Green Version]
  215. Paoletti, P.; Vila, I.; Rifé, M.; Lizcano, J.M.; Alberch, J.; Ginés, S. Dopaminergic and Glutamatergic Signaling Crosstalk in Huntington’s Disease Neurodegeneration: The Role of p25/Cyclin-Dependent Kinase 5. J. Neurosci. 2008, 28, 10090–10101. [Google Scholar] [CrossRef] [Green Version]
  216. Brito, V.; Giralt, A.; Masana, M.; Royes, A.; Espina, M.; Sieiro, E.; Alberch, J.; Castañé, A.; Girault, J.A.; Ginés, S. Cyclin-Dependent Kinase 5 Dysfunction Contributes to Depressive-like Behaviors in Huntington’s Disease by Altering the DARPP-32 Phosphorylation Status in the Nucleus Accumbens. Biol. Psychiatry 2019, 86, 196–207. [Google Scholar] [CrossRef] [PubMed]
  217. Ko, Y.U.; Kim, C.; Lee, J.; Kim, D.; Kim, Y.; Yun, N.; Oh, Y.J. Site-specific phosphorylation of Fbxw7 by Cdk5/p25 and its resulting decreased stability are linked to glutamate-induced excitotoxicity. Cell Death Dis. 2019, 10, 579. [Google Scholar] [CrossRef]
  218. Maestre, C.; Delgado-Esteban, M.; Gomez Sanchez, J.C.; Bolaños, J.P.; Almeida, A. Cdk5 phosphorylates Cdh1 and modulates cyclin B1 stability in excitotoxicity. EMBO J. 2008, 27, 2736–2745. [Google Scholar] [CrossRef] [Green Version]
  219. Damri, O.; Shemesh, N.; Agam, G. Is There Justification to Treat Neurodegenerative Disorders by Repurposing Drugs? The Case of Alzheimer’s Disease, Lithium, and Autophagy. Int. J. Mol. Sci. 2020, 22, 189. [Google Scholar] [CrossRef]
  220. 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] [Green Version]
  221. Caccamo, A.; Magrì, A.; Medina, D.X.; Wisely, E.V.; López Aranda, M.F.; Silva, A.J.; Oddo, S. mTOR regulates tau phosphorylation and degradation: Implications for Alzheimer’s disease and other tauopathies. Aging Cell 2013, 12, 370–380. [Google Scholar] [CrossRef] [Green Version]
  222. Avila, J.; Wandosell, F.; Hernández, F. Role of glycogen synthase kinase-3 in Alzheimer’s disease pathogenesis and glycogen synthase kinase-3 inhibitors. Expert Rev. Neurother. 2010, 10, 703–710. [Google Scholar] [CrossRef] [PubMed]
  223. Llorens-Martín, M.; Jurado, J.; Hernández, F.; Avila, J. GSK-3β, a pivotal kinase in Alzheimer disease. Front. Mol. Neurosci. 2014, 7, 46. [Google Scholar] [PubMed] [Green Version]
  224. Takashima, A. GSK-3 is essential in the pathogenesis of Alzheimer’s disease. J. Alzheimers Dis. 2006, 9 (Suppl. 3), 309–317. [Google Scholar] [CrossRef]
  225. Malhi, G.S.; Outhred, T. Therapeutic Mechanisms of Lithium in Bipolar Disorder: Recent Advances and Current Understanding. CNS Drugs 2016, 30, 931–949. [Google Scholar] [CrossRef] [PubMed]
  226. O’Brien, W.T.; Klein, P.S. Validating GSK3 as an in vivo target of lithium action. Biochem. Soc. Trans. 2009, 37 Pt 5, 133–138. [Google Scholar]
  227. Nunes, M.A.; Schöwe, N.M.; Monteiro-Silva, K.C.; Baraldi-Tornisielo, T.; Souza, S.I.G.; Balthazar, J.; Albuquerque, M.S.; Caetano, A.L.; Viel, T.A.; Buck, H.S. Chronic Microdose Lithium Treatment Prevented Memory Loss and Neurohistopathological Changes in a Transgenic Mouse Model of Alzheimer’s Disease. PLoS ONE 2015, 10, e0142267. [Google Scholar]
  228. Rockenstein, E.; Torrance, M.; Adame, A.; Mante, M.; Bar-On, P.; Rose, J.B.; Crews, L.; Masliah, E. Neuroprotective effects of regulators of the glycogen synthase kinase-3beta signaling pathway in a transgenic model of Alzheimer’s disease are associated with reduced amyloid precursor protein phosphorylation. J. Neurosci. 2007, 27, 1981–1991. [Google Scholar] [CrossRef]
  229. Noble, W.; Planel, E.; Zehr, C.; Olm, V.; Meyerson, J.; Suleman, F.; Gaynor, K.; Wang, L.; LaFrancois, J.; Feinstein, B.; et al. Inhibition of glycogen synthase kinase-3 by lithium correlates with reduced tauopathy and degeneration in vivo. Proc. Natl. Acad. Sci. USA 2005, 102, 6990–6995. [Google Scholar] [CrossRef] [Green Version]
  230. Sofola, O.; Kerr, F.; Rogers, I.; Killick, R.; Augustin, H.; Gandy, C.; Allen, M.J.; Hardy, J.; Lovestone, S.; Partridge, L. Inhibition of GSK-3 ameliorates Abeta pathology in an adult-onset Drosophila model of Alzheimer’s disease. PLoS Genet. 2010, 6, e1001087. [Google Scholar] [CrossRef] [Green Version]
  231. Adityanjee null Munshi, K.R.; Thampy, A. The syndrome of irreversible lithium-effectuated neurotoxicity. Clin. Neuropharmacol. 2005, 28, 38–49. [Google Scholar] [CrossRef] [PubMed]
  232. Sheean, G.L. Lithium neurotoxicity. Clin. Exp. Neurol. 1991, 28, 112–127. [Google Scholar]
  233. Sun, M.; Herrmann, N.; Shulman, K.I. Lithium Toxicity in Older Adults: A Systematic Review of Case Reports. Clin. Drug Investig. 2018, 38, 201–209. [Google Scholar] [CrossRef]
  234. Roux, M.; Dosseto, A. From direct to indirect lithium targets: A comprehensive review of omics data. Metallomics 2017, 9, 1326–1351. [Google Scholar] [CrossRef] [PubMed]
  235. Agranoff, B.W.; Fisher, S.K. Inositol, lithium, and the Brain. Psychopharmacol. Bull. 2001, 35, 5–18. [Google Scholar]
  236. Forlenza, O.V.; Diniz, B.S.; Radanovic, M.; Santos, F.S.; Talib, L.L.; Gattaz, W.F. Disease-modifying properties of long-term lithium treatment for amnestic mild cognitive impairment: Randomised controlled trial. Br. J. Psychiatry 2011, 198, 351–356. [Google Scholar] [CrossRef] [PubMed]
  237. Nunes, M.A.; Viel, T.A.; Buck, H.S. Microdose lithium treatment stabilized cognitive impairment in patients with Alzheimer’s disease. Curr. Alzheimer Res. 2013, 10, 104–107. [Google Scholar]
  238. Dunn, N.; Holmes, C.; Mullee, M. Does lithium therapy protect against the onset of dementia? Alzheimer Dis. Assoc. Disord. 2005, 19, 20–22. [Google Scholar] [CrossRef]
  239. Kessing, L.V.; Søndergård, L.; Forman, J.L.; Andersen, P.K. Lithium treatment and risk of dementia. Arch. Gen. Psychiatry 2008, 65, 1331–1335. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  240. Macdonald, A.; Briggs, K.; Poppe, M.; Higgins, A.; Velayudhan, L.; Lovestone, S. A feasibility and tolerability study of lithium in Alzheimer’s disease. Int. J. Geriatr. Psychiatry 2008, 23, 704–711. [Google Scholar] [CrossRef] [PubMed]
  241. Del Ser, T.; Steinwachs, K.C.; Gertz, H.J.; Andres, M.V.; Gomez-Carrillo, B.; Medina, M.; Vericat, J.A.; Redondo, P.; Fleet, D.; Leon, T. Treatment of Alzheimer’s disease with the GSK-3 inhibitor tideglusib: A pilot study. J. Alzheimers Dis. 2013, 33, 205–215. [Google Scholar] [CrossRef] [PubMed]
  242. Lovestone, S.; Boada, M.; Dubois, B.; Hüll, M.; Rinne, J.O.; Huppertz, H.J.; Calero, M.; Andres, M.V.; Gómez-Carrillo, B.; Leon, T.; et al. A phase II trial of tideglusib in Alzheimer’s disease. J. Alzheimers Dis. 2015, 45, 75–88. [Google Scholar] [CrossRef] [PubMed]
  243. Bhat, R.V.; Andersson, U.; Andersson, S.; Knerr, L.; Bauer, U.; Sundgren-Andersson, A.K. The Conundrum of GSK3 Inhibitors: Is it the Dawn of a New Beginning? J. Alzheimers Dis. 2018, 64, S547–S554. [Google Scholar] [CrossRef] [PubMed]
  244. Georgievska, B.; Sandin, J.; Doherty, J.; Mörtberg, A.; Neelissen, J.; Andersson, A.; Gruber, S.; Nilsson, Y.; Schött, P.; Arvidsson, P.I.; et al. AZD1080, a novel GSK3 inhibitor, rescues synaptic plasticity deficits in rodent brain and exhibits peripheral target engagement in humans. J. Neurochem. 2013, 125, 446–456. [Google Scholar] [CrossRef] [PubMed]
  245. Bhatia, K.; Bhumika null Das, A. Combinatorial drug therapy in cancer—New insights. Life Sci. 2020, 258, 118134. [Google Scholar] [CrossRef]
  246. Al-Lazikani, B.; Banerji, U.; Workman, P. Combinatorial drug therapy for cancer in the post-genomic era. Nat. Biotechnol. 2012, 30, 679–692. [Google Scholar] [CrossRef] [PubMed]
  247. Yao, H.; Wang, H.; Li, C.; Fang, J.-Y.; Xu, J. Cancer Cell-Intrinsic PD-1 and Implications in Combinatorial Immunotherapy. Front. Immunol. 2018, 9, 1774. [Google Scholar] [CrossRef]
  248. Griebel, G.; Stemmelin, J.; Lopez-Grancha, M.; Boulay, D.; Boquet, G.; Slowinski, F.; Pichat, P.; Beeské, S.; Tanaka, S.; Mori, A.; et al. The selective GSK3 inhibitor, SAR502250, displays neuroprotective activity and attenuates behavioral impairments in models of neuropsychiatric symptoms of Alzheimer’s disease in rodents. Sci. Rep. 2019, 9, 18045. [Google Scholar] [CrossRef] [Green Version]
  249. Bhat, R.; Xue, Y.; Berg, S.; Hellberg, S.; Ormö, M.; Nilsson, Y.; Radesäter, A.C.; Jerning, E.; Markgren, P.O.; Borgegård, T.; et al. Structural insights and biological effects of glycogen synthase kinase 3-specific inhibitor AR-A014418. J. Biol. Chem. 2003, 278, 45937–45945. [Google Scholar] [CrossRef] [Green Version]
  250. Mudher, A.; Shepherd, D.; Newman, T.A.; Mildren, P.; Jukes, J.P.; Squire, A.; Mears, A.; Berg, S.; MacKay, D.; Asuni, A.A.; et al. GSK-3beta inhibition reverses axonal transport defects and behavioural phenotypes in Drosophila. Mol. Psychiatry 2004, 9, 522–530. [Google Scholar] [CrossRef] [Green Version]
  251. Selenica, M.L.; Jensen, H.S.; Larsen, A.K.; Pedersen, M.L.; Helboe, L.; Leist, M.; Lotharius, J. Efficacy of small-molecule glycogen synthase kinase-3 inhibitors in the postnatal rat model of tau hyperphosphorylation. Br. J. Pharmacol. 2007, 152, 959–979. [Google Scholar] [CrossRef] [Green Version]
  252. Huang, H.J.; Chen, S.L.; Huang, H.Y.; Sun, Y.C.; Lee, G.C.; Lee-Chen, G.J.; Hsieh-Li, H.M.; Su, M.T. Chronic low dose of AM404 ameliorates the cognitive impairment and pathological features in hyperglycemic 3xTg-AD mice. Psychopharmacology 2019, 236, 763–773. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  253. Abush, H.; Akirav, I. Cannabinoids modulate hippocampal memory and plasticity. Hippocampus 2010, 20, 1126–1138. [Google Scholar] [CrossRef] [PubMed]
  254. Luo, G.; Chen, L.; Burton, C.R.; Xiao, H.; Sivaprakasam, P.; Krause, C.M.; Cao, Y.; Liu, N.; Lippy, J.; Clarke, W.J.; et al. Discovery of Isonicotinamides as Highly Selective, Brain Penetrable, and Orally Active Glycogen Synthase Kinase-3 Inhibitors. J. Med. Chem. 2016, 59, 1041–1051. [Google Scholar] [CrossRef]
  255. Hu, S.; Begum, A.N.; Jones, M.R.; Oh, M.S.; Beech, W.K.; Beech, B.H.; Yang, F.; Chen, P.; Ubeda, O.J.; Kim, P.C.; et al. GSK3 inhibitors show benefits in an Alzheimer’s disease (AD) model of neurodegeneration but adverse effects in control animals. Neurobiol. Dis. 2009, 33, 193–206. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Akiguchi, I.; Pallàs, M.; Budka, H.; Akiyama, H.; Ueno, M.; Han, J.; Yagi, H.; Nishikawa, T.; Chiba, Y.; Sugiyama, H.; et al. SAMP8 mice as a neuropathological model of accelerated brain aging and dementia: Toshio Takeda’s legacy and future directions. Neuropathology 2017, 37, 293–305. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Canudas, A.M.; Gutierrez-Cuesta, J.; Rodríguez, M.I.; Acuña-Castroviejo, D.; Sureda, F.X.; Camins, A.; Pallàs, M. Hyperphosphorylation of microtubule-associated protein tau in senescence-accelerated mouse (SAM). Mech. Ageing Dev. 2005, 126, 1300–1304. [Google Scholar] [CrossRef]
  258. Valencia, A.; Reeves, P.B.; Sapp, E.; Li, X.; Alexander, J.; Kegel, K.B.; Chase, K.; Aronin, N.; DiFiglia, M. Mutant huntingtin and glycogen synthase kinase 3-beta accumulate in neuronal lipid rafts of a presymptomatic knock-in mouse model of Huntington’s disease. J. Neurosci. Res. 2010, 88, 179–190. [Google Scholar] [CrossRef]
  259. Rosas, H.D.; Koroshetz, W.J.; Chen, Y.I.; Skeuse, C.; Vangel, M.; Cudkowicz, M.E.; Caplan, K.; Marek, K.; Seidman, L.J.; Makris, N.; et al. Evidence for more widespread cerebral pathology in early HD: An MRI-based morphometric analysis. Neurology 2003, 60, 1615–1620. [Google Scholar] [CrossRef]
  260. Ransome, M.I.; Hannan, A.J. Behavioural state differentially engages septohippocampal cholinergic and GABAergic neurons in R6/1 Huntington’s disease mice. Neurobiol. Learn Mem. 2012, 97, 261–270. [Google Scholar] [CrossRef]
  261. Ransome, M.I.; Renoir, T.; Hannan, A.J. Hippocampal neurogenesis, cognitive deficits and affective disorder in Huntington’s disease. Neural Plast. 2012, 2012, 874387. [Google Scholar] [CrossRef]
  262. Spargo, E.; Everall, I.P.; Lantos, P.L. Neuronal loss in the hippocampus in Huntington’s disease: A comparison with HIV infection. J. Neurol. Neurosurg. Psychiatry 1993, 56, 487–491. [Google Scholar] [CrossRef]
  263. Humbert, S.; Bryson, E.A.; Cordelières, F.P.; Connors, N.C.; Datta, S.R.; Finkbeiner, S.; Greenberg, M.E.; Saudou, F. The IGF-1/Akt pathway is neuroprotective in Huntington’s disease and involves Huntingtin phosphorylation by Akt. Dev. Cell 2002, 2, 831–837. [Google Scholar] [CrossRef] [Green Version]
  264. Emamian, E.S. AKT/GSK3 signaling pathway and schizophrenia. Front. Mol. Neurosci. 2012, 5, 33. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  265. Zhang, Y.; Huang, N.Q.; Yan, F.; Jin, H.; Zhou, S.Y.; Shi, J.S.; Jin, F. Diabetes mellitus and Alzheimer’s disease: GSK-3β as a potential link. Behav. Brain Res. 2018, 339, 57–65. [Google Scholar] [CrossRef]
  266. Carmichael, J.; Sugars, K.L.; Bao, Y.P.; Rubinsztein, D.C. Glycogen synthase kinase-3beta inhibitors prevent cellular polyglutamine toxicity caused by the Huntington’s disease mutation. J. Biol. Chem. 2002, 277, 33791–33798. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  267. Voisine, C.; Varma, H.; Walker, N.; Bates, E.A.; Stockwell, B.R.; Hart, A.C. Identification of potential therapeutic drugs for huntington’s disease using Caenorhabditis elegans. PLoS ONE 2007, 2, e504. [Google Scholar] [CrossRef] [PubMed]
  268. Wood, N.I.; Morton, A.J. Chronic lithium chloride treatment has variable effects on motor behaviour and survival of mice transgenic for the Huntington’s disease mutation. Brain Res. Bull. 2003, 61, 375–383. [Google Scholar] [CrossRef]
  269. Chiu, C.-T.; Liu, G.; Leeds, P.; Chuang, D.-M. Combined treatment with the mood stabilizers lithium and valproate produces multiple beneficial effects in transgenic mouse models of Huntington’s disease. Neuropsychopharmacology 2011, 36, 2406–2421. [Google Scholar] [CrossRef] [Green Version]
  270. Romoli, M.; Mazzocchetti, P.; D’Alonzo, R.; Siliquini, S.; Rinaldi, V.E.; Verrotti, A.; Calabresi, P.; Costa, C. Valproic Acid and Epilepsy: From Molecular Mechanisms to Clinical Evidences. Curr. Neuropharmacol. 2019, 17, 926–946. [Google Scholar] [CrossRef]
  271. Ximenes, J.C.M.; Neves, K.R.T.; Leal, L.K.A.; do Carmo, M.R.S.; Brito, G.A.D.C.; Naffah-Mazzacoratti, M.D.G.; Cavalheiro, É.A.; Viana, G.S.D.B. Valproic Acid Neuroprotection in the 6-OHDA Model of Parkinson’s Disease Is Possibly Related to Its Anti-Inflammatory and HDAC Inhibitory Properties. J. Neurodegener. Dis. 2015, 2015, 313702. [Google Scholar] [CrossRef]
  272. Sarkar, S.; Krishna, G.; Imarisio, S.; Saiki, S.; O’Kane, C.J.; Rubinsztein, D.C. A rational mechanism for combination treatment of Huntington’s disease using lithium and rapamycin. Hum. Mol. Genet. 2008, 17, 170–178. [Google Scholar] [CrossRef] [Green Version]
  273. Fernández-Nogales, M.; Hernández, F.; Miguez, A.; Alberch, J.; Ginés, S.; Pérez-Navarro, E.; Lucas, J.J. Decreased glycogen synthase kinase-3 levels and activity contribute to Huntington’s disease. Hum. Mol. Genet. 2015, 24, 5040–5052. [Google Scholar] [CrossRef] [PubMed]
  274. Rai, A.; Singh, P.K.; Singh, V.; Kumar, V.; Mishra, R.; Thakur, A.K.; Mahadevan, A.; Shankar, S.K.; Jana, N.R.; Ganesh, S. Glycogen synthase protects neurons from cytotoxicity of mutant huntingtin by enhancing the autophagy flux. Cell Death Dis. 2018, 9, 201. [Google Scholar] [CrossRef] [PubMed]
  275. Cargnello, M.; Roux, P.P. Activation and function of the MAPKs and their substrates, the MAPK-activated protein kinases. Microbiol. Mol. Biol. Rev. 2011, 75, 50–83. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  276. Cuenda, A.; Rousseau, S. p38 MAP-kinases pathway regulation, function and role in human diseases. Biochim. Biophys. Acta 2007, 1773, 1358–1375. [Google Scholar] [CrossRef] [Green Version]
  277. Nebreda, A.R.; Porras, A. p38 MAP kinases: Beyond the stress response. Trends Biochem. Sci. 2000, 25, 257–260. [Google Scholar] [CrossRef]
  278. Ashwell, J.D. The many paths to p38 mitogen-activated protein kinase activation in the immune system. Nat. Rev. Immunol. 2006, 6, 532–540. [Google Scholar] [CrossRef]
  279. Martínez-Limón, A.; Joaquin, M.; Caballero, M.; Posas, F.; de Nadal, E. The p38 Pathway: From Biology to Cancer Therapy. Int. J. Mol. Sci. 2020, 21, 1913. [Google Scholar] [CrossRef] [Green Version]
  280. Aouadi, M.; Binetruy, B.; Caron, L.; Le Marchand-Brustel, Y.; Bost, F. Role of MAPKs in development and differentiation: Lessons from knockout mice. Biochimie 2006, 88, 1091–1098. [Google Scholar] [CrossRef] [PubMed]
  281. Coulthard, L.R.; White, D.E.; Jones, D.L.; McDermott, M.F.; Burchill, S.A. p38(MAPK): Stress responses from molecular mechanisms to therapeutics. Trends Mol. Med. 2009, 15, 369–379. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  282. Zarubin, T.; Han, J. Activation and signaling of the p38 MAP kinase pathway. Cell Res. 2005, 15, 11–18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  283. Asih, P.R.; Prikas, E.; Stefanoska, K.; Tan, A.R.P.; Ahel, H.I.; Ittner, A. Functions of p38 MAP Kinases in the Central Nervous System. Front. Mol. Neurosci. 2020, 13, 570586. [Google Scholar] [CrossRef] [PubMed]
  284. Mittelstadt, P.R.; Salvador, J.M.; Fornace, A.J.; Ashwell, J.D. Activating p38 MAPK: New tricks for an old kinase. Cell Cycle 2005, 4, 1189–1192. [Google Scholar] [CrossRef] [Green Version]
  285. Raman, M.; Earnest, S.; Zhang, K.; Zhao, Y.; Cobb, M.H. TAO kinases mediate activation of p38 in response to DNA damage. EMBO J. 2007, 26, 2005–2014. [Google Scholar] [CrossRef]
  286. Raman, M.; Chen, W.; Cobb, M.H. Differential regulation and properties of MAPKs. Oncogene 2007, 26, 3100–3112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  287. Corrêa, S.A.L.; Eales, K.L. The Role of p38 MAPK and Its Substrates in Neuronal Plasticity and Neurodegenerative Disease. J. Signal Transduct. 2012, 2012, 649079. [Google Scholar] [CrossRef] [Green Version]
  288. Segarra, J.; Balenci, L.; Drenth, T.; Maina, F.; Lamballe, F. Combined signaling through ERK, PI3K/AKT, and RAC1/p38 is required for met-triggered cortical neuron migration. J. Biol. Chem. 2006, 281, 4771–4778. [Google Scholar] [CrossRef] [Green Version]
  289. Zhen, X.; Du, W.; Romano, A.G.; Friedman, E.; Harvey, J.A. The p38 mitogen-activated protein kinase is involved in associative learning in rabbits. J. Neurosci. 2001, 21, 5513–5519. [Google Scholar] [CrossRef] [Green Version]
  290. Grinkevich, L.N. p38 MAPK is Involved in Regulation of Epigenetic Mechanisms of Food Aversion Learning. Bull. Exp. Biol. Med. 2017, 163, 412–414. [Google Scholar] [CrossRef]
  291. Navarrete, M.; Cuartero, M.I.; Palenzuela, R.; Draffin, J.E.; Konomi, A.; Serra, I.; Colié, S.; Castaño-Castaño, S.; Hasan, M.T.; Nebreda, Á.R.; et al. Astrocytic p38α MAPK drives NMDA receptor-dependent long-term depression and modulates long-term memory. Nat. Commun. 2019, 10, 2968. [Google Scholar] [CrossRef] [Green Version]
  292. Bolshakov, V.Y.; Carboni, L.; Cobb, M.H.; Siegelbaum, S.A.; Belardetti, F. Dual MAP kinase pathways mediate opposing forms of long-term plasticity at CA3-CA1 synapses. Nat. Neurosci. 2000, 3, 1107–1112. [Google Scholar] [CrossRef]
  293. Moult, P.R.; Corrêa, S.A.L.; Collingridge, G.L.; Fitzjohn, S.M.; Bashir, Z.I. Co-activation of p38 mitogen-activated protein kinase and protein tyrosine phosphatase underlies metabotropic glutamate receptor-dependent long-term depression. J. Physiol. 2008, 586, 2499–2510. [Google Scholar] [CrossRef]
  294. Hsieh, H.; Boehm, J.; Sato, C.; Iwatsubo, T.; Tomita, T.; Sisodia, S.; Malinow, R. AMPAR removal underlies Abeta-induced synaptic depression and dendritic spine loss. Neuron 2006, 52, 831–843. [Google Scholar] [CrossRef] [Green Version]
  295. Izumi, Y.; Zorumski, C.F. Temperoammonic Stimulation Depotentiates Schaffer Collateral LTP via p38 MAPK Downstream of Adenosine A1 Receptors. J. Neurosci. 2019, 39, 1783–1792. [Google Scholar] [CrossRef]
  296. Izumi, Y.; Tokuda, K.; Zorumski, C.F. Long-term potentiation inhibition by low-level N-methyl-D-aspartate receptor activation involves calcineurin, nitric oxide, and p38 mitogen-activated protein kinase. Hippocampus 2008, 18, 258–265. [Google Scholar] [CrossRef]
  297. Wu, R.; Chen, H.; Ma, J.; He, Q.; Huang, Q.; Liu, Q.; Li, M.; Yuan, Z. c-Abl-p38α signaling plays an important role in MPTP-induced neuronal death. Cell Death Differ. 2016, 23, 542–552. [Google Scholar] [CrossRef] [Green Version]
  298. Xing, B.; Bachstetter, A.D.; Van Eldik, L.J. Microglial p38α MAPK is critical for LPS-induced neuron degeneration, through a mechanism involving TNFα. Mol. Neurodegener. 2011, 6, 84. [Google Scholar] [CrossRef] [Green Version]
  299. Choi, W.S.; Eom, D.S.; Han, B.S.; Kim, W.K.; Han, B.H.; Choi, E.J.; Oh, T.H.; Markelonis, G.J.; Cho, J.W.; Oh, Y.J. Phosphorylation of p38 MAPK induced by oxidative stress is linked to activation of both caspase-8-and-9-mediated apoptotic pathways in dopaminergic neurons. J. Biol. Chem. 2004, 279, 20451–20460. [Google Scholar] [CrossRef] [Green Version]
  300. Fan, C.; Long, Y.; Wang, L.; Liu, X.; Liu, Z.; Lan, T.; Li, Y.; Yu, S.Y. N-Acetylcysteine Rescues Hippocampal Oxidative Stress-Induced Neuronal Injury via Suppression of p38/JNK Signaling in Depressed Rats. Front. Cell Neurosci. 2020, 14, 554613. [Google Scholar] [CrossRef]
  301. Xing, B.; Bachstetter, A.D.; Van Eldik, L.J. Inhibition of neuronal p38α, but not p38β MAPK, provides neuroprotection against three different neurotoxic insults. J. Mol. Neurosci. 2015, 55, 509–518. [Google Scholar] [CrossRef] [Green Version]
  302. Semenova, M.M.; Mäki-Hokkonen, A.M.; Cao, J.; Komarovski, V.; Forsberg, K.M.; Koistinaho, M.; Coffey, E.T.; Courtney, M.J. Rho mediates calcium-dependent activation of p38alpha and subsequent excitotoxic cell death. Nat. Neurosci. 2007, 10, 436–443. [Google Scholar] [CrossRef]
  303. Umezawa, H.; Naito, Y.; Tanaka, K.; Yoshioka, K.; Suzuki, K.; Sudo, T.; Hagihara, M.; Hatano, M.; Tatsumi, K.; Kasuya, Y. Genetic and Pharmacological Inhibition of p38α Improves Locomotor Recovery after Spinal Cord Injury. Front. Pharmacol. 2017, 8, 72. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  304. Coogan, A.N.; O’Neill, L.A.; O’Connor, J.J. The P38 mitogen-activated protein kinase inhibitor SB203580 antagonizes the inhibitory effects of interleukin-1beta on long-term potentiation in the rat dentate gyrus in vitro. Neuroscience 1999, 93, 57–69. [Google Scholar] [CrossRef]
  305. Jha, S.K.; Jha, N.K.; Kar, R.; Ambasta, R.K.; Kumar, P. p38 MAPK and PI3K/AKT Signalling Cascades inParkinson’s Disease. Int. J. Mol. Cell Med. 2015, 4, 67–86. [Google Scholar] [PubMed]
  306. Pei, J.J.; Braak, E.; Braak, H.; Grundke-Iqbal, I.; Iqbal, K.; Winblad, B.; Cowburn, R.F. Localization of active forms of C-jun kinase (JNK) and p38 kinase in Alzheimer’s disease brains at different stages of neurofibrillary degeneration. J. Alzheimers Dis. 2001, 3, 41–48. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  307. Hensley, K.; Floyd, R.A.; Zheng, N.Y.; Nael, R.; Robinson, K.A.; Nguyen, X.; Pye, Q.N.; Stewart, C.A.; Geddes, J.; Markesbery, W.R.; et al. p38 kinase is activated in the Alzheimer’s disease brain. J. Neurochem. 1999, 72, 2053–2058. [Google Scholar] [CrossRef] [PubMed]
  308. Rottkamp, C.A.; Raina, A.K.; Zhu, X.; Gaier, E.; Bush, A.I.; Atwood, C.S.; Chevion, M.; Perry, G.; Smith, M.A. Redox-active iron mediates amyloid-beta toxicity. Free Radic. Biol. Med. 2001, 30, 447–450. [Google Scholar] [CrossRef]
  309. Bhaskar, K.; Konerth, M.; Kokiko-Cochran, O.N.; Cardona, A.; Ransohoff, R.M.; Lamb, B.T. Regulation of tau pathology by the microglial fractalkine receptor. Neuron 2010, 68, 19–31. [Google Scholar] [CrossRef] [Green Version]
  310. Zhu, X.; Rottkamp, C.A.; Hartzler, A.; Sun, Z.; Takeda, A.; Boux, H.; Shimohama, S.; Perry, G.; Smith, M.A. Activation of MKK6, an upstream activator of p38, in Alzheimer’s disease. J. Neurochem. 2001, 79, 311–318. [Google Scholar] [CrossRef]
  311. Giovannini, M.G.; Scali, C.; Prosperi, C.; Bellucci, A.; Vannucchi, M.G.; Rosi, S.; Pepeu, G.; Casamenti, F. Beta-amyloid-induced inflammation and cholinergic hypofunction in the rat brain in vivo: Involvement of the p38MAPK pathway. Neurobiol. Dis. 2002, 11, 257–274. [Google Scholar] [CrossRef] [Green Version]
  312. Bodles, A.M.; Barger, S.W. Secreted beta-amyloid precursor protein activates microglia via JNK and p38-MAPK. Neurobiol. Aging 2005, 26, 9–16. [Google Scholar] [CrossRef]
  313. Culbert, A.A.; Skaper, S.D.; Howlett, D.R.; Evans, N.A.; Facci, L.; Soden, P.E.; Seymour, Z.M.; Guillot, F.; Gaestel, M.; Richardson, J.C.; et al. MAPK-activated protein kinase 2 deficiency in microglia inhibits pro-inflammatory mediator release and resultant neurotoxicity. Relevance to neuroinflammation in a transgenic mouse model of Alzheimer disease. J. Biol. Chem. 2006, 281, 23658–23667. [Google Scholar] [CrossRef] [Green Version]
  314. Munoz, L.; Ranaivo, H.R.; Roy, S.M.; Hu, W.; Craft, J.M.; McNamara, L.K.; Chico, L.W.; Van Eldik, L.J.; Watterson, D.M. A novel p38 alpha MAPK inhibitor suppresses brain proinflammatory cytokine up-regulation and attenuates synaptic dysfunction and behavioral deficits in an Alzheimer’s disease mouse model. J. Neuroinflamm. 2007, 4, 21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  315. Ko, J.H.; Yoon, S.-O.; Lee, H.J.; Oh, J.Y. Rapamycin regulates macrophage activation by inhibiting NLRP3 inflammasome-p38 MAPK-NFκB pathways in autophagy- and p62-dependent manners. Oncotarget 2017, 8, 40817–40831. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  316. Ittner, A.; Chua, S.W.; Bertz, J.; Volkerling, A.; van der Hoven, J.; Gladbach, A.; Przybyla, M.; Bi, M.; van Hummel, A.; Stevens, C.H.; et al. Site-specific phosphorylation of tau inhibits amyloid-β toxicity in Alzheimer’s mice. Science 2016, 354, 904–908. [Google Scholar] [CrossRef] [Green Version]
  317. Ittner, A.; Asih, P.R.; Tan, A.R.; Prikas, E.; Bertz, J.; Stefanoska, K.; Lin, Y.; Volkerling, A.M.; Ke, Y.D.; Delerue, F.; et al. Reduction of advanced tau-mediated memory deficits by the MAP kinase p38γ. Acta Neuropathol. 2020, 140, 279–294. [Google Scholar] [CrossRef]
  318. Yoshida, H.; Goedert, M. Sequential phosphorylation of tau protein by cAMP-dependent protein kinase and SAPK4/p38delta or JNK2 in the presence of heparin generates the AT100 epitope. J. Neurochem. 2006, 99, 154–164. [Google Scholar] [CrossRef]
  319. Buée-Scherrer, V.; Goedert, M. Phosphorylation of microtubule-associated protein tau by stress-activated protein kinases in intact Cells. FEBS Lett. 2002, 515, 151–154. [Google Scholar] [CrossRef] [Green Version]
  320. Roy, S.M.; Grum-Tokars, V.L.; Schavocky, J.P.; Saeed, F.; Staniszewski, A.; Teich, A.F.; Arancio, O.; Bachstetter, A.D.; Webster, S.J.; Van Eldik, L.J.; et al. Targeting human central nervous system protein kinases: An isoform selective p38αMAPK inhibitor that attenuates disease progression in Alzheimer’s disease mouse models. ACS Chem. Neurosci. 2015, 6, 666–680. [Google Scholar] [CrossRef] [Green Version]
  321. Zhou, Z.; Bachstetter, A.D.; Späni, C.B.; Roy, S.M.; Watterson, D.M.; Van Eldik, L.J. Retention of normal glia function by an isoform-selective protein kinase inhibitor drug candidate that modulates cytokine production and cognitive outcomes. J. Neuroinflamm. 2017, 14, 75. [Google Scholar] [CrossRef] [Green Version]
  322. Alam, J.J. Selective Brain-Targeted Antagonism of p38 MAPKα Reduces Hippocampal IL-1β Levels and Improves Morris Water Maze Performance in Aged Rats. J. Alzheimers Dis. 2015, 48, 219–227. [Google Scholar] [CrossRef] [Green Version]
  323. Gee, M.S.; Son, S.H.; Jeon, S.H.; Do, J.; Kim, N.; Ju, Y.J.; Lee, S.J.; Chung, E.K.; Inn, K.S.; Kim, N.J.; et al. A selective p38α/β MAPK inhibitor alleviates neuropathology and cognitive impairment, and modulates microglia function in 5XFAD mouse. Alzheimers Res. Ther. 2020, 12, 45. [Google Scholar] [CrossRef] [Green Version]
  324. Ashabi, G.; Alamdary, S.Z.; Ramin, M.; Khodagholi, F. Reduction of hippocampal apoptosis by intracerebroventricular administration of extracellular signal-regulated protein kinase and/or p38 inhibitors in amyloid beta rat model of Alzheimer’s disease: Involvement of nuclear-related factor-2 and nuclear factor-κB. Basic Clin. Pharmacol. Toxicol. 2013, 112, 145–155. [Google Scholar] [PubMed]
  325. Cui, Y.Q.; Wang, Q.; Zhang, D.M.; Wang, J.Y.; Xiao, B.; Zheng, Y.; Wang, X.M. Triptolide Rescues Spatial Memory Deficits and Amyloid-β Aggregation Accompanied by Inhibition of Inflammatory Responses and MAPKs Activity in APP/PS1 Transgenic Mice. Curr. Alzheimer Res. 2016, 13, 288–296. [Google Scholar] [CrossRef] [PubMed]
  326. Jiang, J.; Wang, Z.; Liang, X.; Nie, Y.; Chang, X.; Xue, H.; Li, S.; Min, C. Intranasal MMI-0100 Attenuates Aβ1-42- and LPS-Induced Neuroinflammation and Memory Impairments via the MK2 Signaling Pathway. Front. Immunol. 2019, 10, 2707. [Google Scholar] [CrossRef]
  327. Muraleva, N.A.; Stefanova, N.A.; Kolosova, N.G. SkQ1 Suppresses the p38 MAPK Signaling Pathway Involved in Alzheimer’s Disease-Like Pathology in OXYS Rats. Antioxidants 2020, 9, 676. [Google Scholar] [CrossRef]
  328. Liang, Z.; Zhang, B.; Xu, M.; Morisseau, C.; Hwang, S.H.; Hammock, B.D.; Qing, X.L. 1-Trifluoromethoxyphenyl-3-(1-propionylpiperidin-4-yl) Urea, a Selective and Potent Dual Inhibitor of Soluble Epoxide Hydrolase and p38 Kinase Intervenes in Alzheimer’s Signaling in Human Nerve Cells. ACS Chem. Neurosci. 2019, 10, 4018–4030. [Google Scholar] [CrossRef]
  329. Wang, Q.; Jiang, H.; Wang, L.; Yi, H.; Li, Z.; Liu, R. Vitegnoside Mitigates Neuronal Injury, Mitochondrial Apoptosis, and Inflammation in an Alzheimer’s Disease Cell Model via the p38 MAPK/JNK Pathway. J. Alzheimers Dis. 2019, 72, 199–214. [Google Scholar] [CrossRef]
  330. El-Din, S.S.; Rashed, L.; Medhat, E.; Aboulhoda, B.E.; Badawy, A.D.; ShamsEldeen, A.M.; Abdelgwad, M. Active form of vitamin D analogue mitigates neurodegenerative changes in Alzheimer’s disease in rats by targeting Keap1/Nrf2 and MAPK-38p/ERK signaling pathways. Steroids 2020, 156, 108586. [Google Scholar] [CrossRef]
  331. Chiu, Y.J.; Hsieh, Y.H.; Lin, T.H.; Lee, G.C.; Hsieh-Li, H.M.; Sun, Y.C.; Chen, C.M.; Chang, K.H.; Lee-Chen, G.J. Novel compound VB-037 inhibits Aβ aggregation and promotes neurite outgrowth through enhancement of HSP27 and reduction of P38 and JNK-mediated inflammation in cell models for Alzheimer’s disease. Neurochem. Int. 2019, 125, 175–186. [Google Scholar] [CrossRef]
  332. Casadomé-Perales, Á.; Matteis, L.D.; Alleva, M.; Infantes-Rodríguez, C.; Palomares-Pérez, I.; Saito, T.; Saido, T.C.; Esteban, J.A.; Nebreda, A.R.; de la Fuente, J.M.; et al. Inhibition of p38 MAPK in the brain through nasal administration of p38 inhibitor loaded in chitosan nanocapsules. Nanomedicine 2019, 14, 2409–2422. [Google Scholar] [CrossRef]
  333. Gianfriddo, M.; Melani, A.; Turchi, D.; Giovannini, M.G.; Pedata, F. Adenosine and glutamate extracellular concentrations and mitogen-activated protein kinases in the striatum of Huntington transgenic mice. Selective antagonism of adenosine A2A receptors reduces transmitter outflow. Neurobiol. Dis. 2004, 17, 77–88. [Google Scholar] [CrossRef] [PubMed]
  334. Rangone, H.; Poizat, G.; Troncoso, J.; Ross, C.A.; MacDonald, M.E.; Saudou, F.; Humbert, S. The serum- and glucocorticoid-induced kinase SGK inhibits mutant huntingtin-induced toxicity by phosphorylating serine 421 of huntingtin. Eur. J. Neurosci. 2004, 19, 273–279. [Google Scholar] [CrossRef]
  335. Reijonen, S.; Kukkonen, J.P.; Hyrskyluoto, A.; Kivinen, J.; Kairisalo, M.; Takei, N.; Lindholm, D.; Korhonen, L. Downregulation of NF-kappaB signaling by mutant huntingtin proteins induces oxidative stress and cell death. Cell Mol. Life Sci. 2010, 67, 1929–1941. [Google Scholar] [PubMed]
  336. Taylor, D.M.; Moser, R.; Régulier, E.; Breuillaud, L.; Dixon, M.; Beesen, A.A.; Elliston, L.; Santos, M.D.F.S.; Kim, J.; Jones, L.; et al. MAP kinase phosphatase 1 (MKP-1/DUSP1) is neuroprotective in Huntington’s disease via additive effects of JNK and p38 inhibition. J. Neurosci. 2013, 33, 2313–2325. [Google Scholar] [CrossRef] [Green Version]
  337. Gladding, C.M.; Fan, J.; Zhang, L.Y.; Wang, L.; Xu, J.; Li, E.H.; Lombroso, P.J.; Raymond, L.A. Alterations in STriatal-Enriched protein tyrosine Phosphatase expression, activation, and downstream signaling in early and late stages of the YAC128 Huntington’s disease mouse model. J. Neurochem. 2014, 130, 145–159. [Google Scholar] [CrossRef] [Green Version]
  338. Pardo, R.; Colin, E.; Régulier, E.; Aebischer, P.; Déglon, N.; Humbert, S.; Saudou, F. Inhibition of calcineurin by FK506 protects against polyglutamine-huntingtin toxicity through an increase of huntingtin phosphorylation at S421. J. Neurosci. 2006, 26, 1635–1645. [Google Scholar] [CrossRef] [Green Version]
  339. Kratter, I.H.; Zahed, H.; Lau, A.; Tsvetkov, A.S.; Daub, A.C.; Weiberth, K.F.; Gu, X.; Saudou, F.; Humbert, S.; Yang, X.W.; et al. Serine 421 regulates mutant huntingtin toxicity and clearance in mice. J. Clin. Investig. 2016, 126, 3585–3597. [Google Scholar] [CrossRef] [Green Version]
  340. Huang, Z.-N.; Chen, J.-M.; Huang, L.-C.; Fang, Y.-H.; Her, L.-S. Inhibition of p38 Mitogen-Activated Protein Kinase Ameliorates HAP40 Depletion-Induced Toxicity and Proteasomal Defect in Huntington’s Disease Model. Mol. Neurobiol. 2021. [Google Scholar] [CrossRef] [PubMed]
  341. Ferrante, A.; Martire, A.; Pepponi, R.; Varani, K.; Vincenzi, F.; Ferraro, L.; Beggiato, S.; Tebano, M.T.; Popoli, P. Expression, pharmacology and functional activity of adenosine A1 receptors in genetic models of Huntington’s disease. Neurobiol. Dis. 2014, 71, 193–204. [Google Scholar] [CrossRef]
  342. Sava, G.P.; Fan, H.; Coombes, R.C.; Buluwela, L.; Ali, S. CDK7 inhibitors as anticancer drugs. Cancer Metastasis Rev. 2020, 39, 805–823. [Google Scholar] [CrossRef]
  343. Poon, R.Y.C. Cell Cycle Control: A System of Interlinking Oscillators. Methods Mol. Biol. 2016, 1342, 3–19. [Google Scholar]
  344. Harashima, H.; Dissmeyer, N.; Schnittger, A. Cell cycle control across the eukaryotic kingdom. Trends Cell Biol. 2013, 23, 345–356. [Google Scholar] [CrossRef] [PubMed]
  345. Malumbres, M.; Harlow, E.; Hunt, T.; Hunter, T.; Lahti, J.M.; Manning, G.; Morgan, D.O.; Tsai, L.H.; Wolgemuth, D.J. Cyclin-dependent kinases: A family portrait. Nat. Cell Biol. 2009, 11, 1275–1276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  346. Lim, S.; Kaldis, P. Cdks, cyclins and CKIs: Roles beyond cell cycle regulation. Development 2013, 140, 3079–3093. [Google Scholar]
  347. Liu, D.X.; Greene, L.A. Neuronal apoptosis at the G1/S cell cycle checkpoint. Cell Tissue Res. 2001, 305, 217–228. [Google Scholar] [CrossRef]
  348. Becker, E.B.E.; Bonni, A. Cell cycle regulation of neuronal apoptosis in development and disease. Prog. Neurobiol. 2004, 72, 1–25. [Google Scholar] [CrossRef] [PubMed]
  349. Bonda, D.J.; Bajić, V.P.; Spremo Potparevic, B.; Casadesus, G.; Zhu, X.; Smith, M.A.; Lee, H.G. Review: Cell cycle aberrations and neurodegeneration. Neuropathol. Appl. Neurobiol. 2010, 36, 157–163. [Google Scholar] [CrossRef]
  350. Aubrecht, T.G.; Faden, A.I.; Sabirzhanov, B.; Glaser, E.P.; Roelofs, B.A.; Polster, B.M.; Makarevich, O.; Stoica, B.A. Comparing effects of CDK inhibition and E2F1/2 ablation on neuronal cell death pathways in vitro and after traumatic brain injury. Cell Death Dis. 2018, 9, 1121. [Google Scholar] [CrossRef] [PubMed]
  351. Strachan, G.D.; Koike, M.A.; Siman, R.; Hall, D.J.; Jordan-Sciutto, K.L. E2F1 induces cell death, calpain activation, and MDMX degradation in a transcription independent manner implicating a novel role for E2F1 in neuronal loss in SIV encephalitis. J. Cell Biochem. 2005, 96, 728–740. [Google Scholar] [CrossRef] [PubMed]
  352. Wang, Y.; Zhou, Y.; Xiao, L.; Zheng, S.; Yan, N.; Chen, D. E2f1 mediates high glucose-induced neuronal death in cultured mouse retinal explants. Cell Cycle 2017, 16, 1824–1834. [Google Scholar] [CrossRef]
  353. Wu, J.; Kharebava, G.; Piao, C.; Stoica, B.A.; Dinizo, M.; Sabirzhanov, B.; Hanscom, M.; Guanciale, K.; Faden, A.I. Inhibition of E2F1/CDK1 pathway attenuates neuronal apoptosis in vitro and confers neuroprotection after spinal cord injury in vivo. PLoS ONE 2012, 7, e42129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  354. Zhang, Y.; Song, X.; Herrup, K. Context-Dependent Functions of E2F1: Cell Cycle, Cell Death, and DNA Damage Repair in Cortical Neurons. Mol. Neurobiol. 2020, 57, 2377–2390. [Google Scholar] [CrossRef] [PubMed]
  355. Verdaguer, E.; de Arriba Susana, G.; Clemens, A.; Pallàs, M.; Camins, A. Implication of the transcription factor E2F-1 in the modulation of neuronal apoptosis. Biomed. Pharmacother. 2007, 61, 390–399. [Google Scholar] [CrossRef] [PubMed]
  356. Yeste-Velasco, M.; Folch, J.; Trullas, R.; Abad, M.A.; Enguita, M.; Pallàs, M.; Camins, A. Glycogen synthase kinase-3 is involved in the regulation of the cell cycle in cerebellar granule Cells. Neuropharmacology 2007, 53, 295–307. [Google Scholar] [CrossRef]
  357. Su, S.C.; Tsai, L.-H. Cyclin-dependent kinases in brain development and disease. Annu. Rev. Cell Dev. Biol. 2011, 27, 465–491. [Google Scholar] [CrossRef] [PubMed]
  358. Kawauchi, T.; Shikanai, M.; Kosodo, Y. Extra-cell cycle regulatory functions of cyclin-dependent kinases (CDK) and CDK inhibitor proteins contribute to brain development and neurological disorders. Genes Cells 2013, 18, 176–194. [Google Scholar] [CrossRef] [Green Version]
  359. Cheffer, A.; Tárnok, A.; Ulrich, H. Cell cycle regulation during neurogenesis in the embryonic and adult Brain. Stem Cell Rev. Rep. 2013, 9, 794–805. [Google Scholar] [CrossRef]
  360. Marlier, Q.; D’aes, T.; Verteneuil, S.; Vandenbosch, R.; Malgrange, B. Core cell cycle machinery is crucially involved in both life and death of post-mitotic neurons. Cell Mol. Life Sci. 2020, 77, 4553–4571. [Google Scholar] [CrossRef]
  361. Namba, T.; Huttner, W.B. Neural progenitor cells and their role in the development and evolutionary expansion of the neocortex. Wiley Interdiscip. Rev. Dev. Biol. 2017. [Google Scholar] [CrossRef]
  362. Namba, T.; Nardelli, J.; Gressens, P.; Huttner, W.B. Metabolic Regulation of Neocortical Expansion in Development and Evolution. Neuron 2020, 109, 408–419. [Google Scholar]
  363. Pinson, A.; Namba, T.; Huttner, W.B. Malformations of Human Neocortex in Development—Their Progenitor Cell Basis and Experimental Model Systems. Front. Cell Neurosci. 2019, 13, 305. [Google Scholar] [CrossRef]
  364. Schmetsdorf, S.; Arnold, E.; Holzer, M.; Arendt, T.; Gärtner, U. A putative role for cell cycle-related proteins in microtubule-based neuroplasticity. Eur. J. Neurosci. 2009, 29, 1096–1107. [Google Scholar] [CrossRef]
  365. Kim, D.; Frank, C.L.; Dobbin, M.M.; Tsunemoto, R.K.; Tu, W.; Peng, P.L.; Guan, J.S.; Lee, B.H.; Moy, L.Y.; Giusti, P.; et al. Deregulation of HDAC1 by p25/Cdk5 in neurotoxicity. Neuron 2008, 60, 803–817. [Google Scholar] [CrossRef] [Green Version]
  366. Herrup, K. Post-mitotic role of the cell cycle machinery. Curr. Opin. Cell Biol. 2013, 25, 711–716. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  367. Löffler, T.; Flunkert, S.; Taub, N.; Schofield, E.L.; Ward, M.A.; Windisch, M.; Hutter-Paier, B. Stable mutated tau441 transfected SH-SY5Y cells as screening tool for Alzheimer’s disease drug candidates. J. Mol. Neurosci. 2012, 47, 192–203. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  368. Lai, K.-O.; Ip, N.Y. Recent advances in understanding the roles of Cdk5 in synaptic plasticity. Biochim. Biophys. Acta 2009, 1792, 741–745. [Google Scholar] [CrossRef] [Green Version]
  369. Angelo, M.; Plattner, F.; Giese, K.P. Cyclin-dependent kinase 5 in synaptic plasticity, learning and memory. J. Neurochem. 2006, 99, 353–370. [Google Scholar] [CrossRef] [PubMed]
  370. Shelton, S.B.; Johnson, G.V.W. Cyclin-dependent kinase-5 in neurodegeneration. J. Neurochem. 2004, 88, 1313–1326. [Google Scholar] [CrossRef] [PubMed]
  371. Ikiz, B.; Przedborski, S. A sequel to the tale of p25/Cdk5 in neurodegeneration. Neuron 2008, 60, 731–732. [Google Scholar] [CrossRef] [Green Version]
  372. Camins, A.; Verdaguer, E.; Folch, J.; Canudas, A.M.; Pallàs, M. The role of CDK5/P25 formation/inhibition in neurodegeneration. Drug News Perspect. 2006, 19, 453–460. [Google Scholar] [CrossRef] [PubMed]
  373. Odajima, J.; Wills, Z.P.; Ndassa, Y.M.; Terunuma, M.; Kretschmannova, K.; Deeb, T.Z.; Geng, Y.; Gawrzak, S.; Quadros, I.M.; Newman, J.; et al. Cyclin E constrains Cdk5 activity to regulate synaptic plasticity and memory formation. Dev. Cell 2011, 21, 655–668. [Google Scholar] [CrossRef] [Green Version]
  374. Varvel, N.H.; Bhaskar, K.; Patil, A.R.; Pimplikar, S.W.; Herrup, K.; Lamb, B.T. Abeta oligomers induce neuronal cell cycle events in Alzheimer’s disease. J. Neurosci. 2008, 28, 10786–10793. [Google Scholar] [CrossRef]
  375. Majd, S.; Zarifkar, A.; Rastegar, K.; Takhshid, M.A. Different fibrillar Abeta 1-42 concentrations induce adult hippocampal neurons to reenter various phases of the Cell Cycle. Brain Res. 2008, 1218, 224–229. [Google Scholar] [CrossRef] [PubMed]
  376. Bhaskar, K.; Miller, M.; Chludzinski, A.; Herrup, K.; Zagorski, M.; Lamb, B.T. The PI3K-Akt-mTOR pathway regulates Abeta oligomer induced neuronal cell cycle events. Mol. Neurodegener. 2009, 4, 14. [Google Scholar] [CrossRef] [Green Version]
  377. Chang, K.-H.; Vincent, F.; Shah, K. Deregulated Cdk5 triggers aberrant activation of cell cycle kinases and phosphatases inducing neuronal death. J. Cell Sci. 2012, 125 Pt 21, 5124–5137. [Google Scholar] [CrossRef] [Green Version]
  378. Pope, W.B.; Lambert, M.P.; Leypold, B.; Seupaul, R.; Sletten, L.; Krafft, G.; Klein, W.L. Microtubule-associated protein tau is hyperphosphorylated during mitosis in the human neuroblastoma cell line SH-SY5Y. Exp. Neurol. 1994, 126, 185–194. [Google Scholar] [CrossRef]
  379. Lindwall, G.; Cole, R.D. Phosphorylation affects the ability of tau protein to promote microtubule assembly. J. Biol. Chem. 1984, 259, 5301–5305. [Google Scholar] [CrossRef]
  380. Webber, K.M.; Raina, A.K.; Marlatt, M.W.; Zhu, X.; Prat, M.I.; Morelli, L.; Casadesus, G.; Perry, G.; Smith, M.A. The cell cycle in Alzheimer disease: A unique target for neuropharmacology. Mech. Ageing Dev. 2005, 126, 1019–1025. [Google Scholar] [CrossRef] [PubMed]
  381. Seward, M.E.; Swanson, E.; Norambuena, A.; Reimann, A.; Cochran, J.N.; Li, R.; Roberson, E.D.; Bloom, G.S. Amyloid-β signals through tau to drive ectopic neuronal cell cycle re-entry in Alzheimer’s disease. J. Cell Sci. 2013, 126 Pt 5, 1278–1286. [Google Scholar] [CrossRef] [Green Version]
  382. Herrup, K.; Neve, R.; Ackerman, S.L.; Copani, A. Divide and die: Cell cycle events as triggers of nerve cell death. J. Neurosci. 2004, 24, 9232–9239. [Google Scholar] [CrossRef] [PubMed]
  383. Song, B.; Davis, K.; Liu, X.S.; Lee, H.; Smith, M.; Liu, X. Inhibition of Polo-like kinase 1 reduces beta-amyloid-induced neuronal cell death in Alzheimer’s disease. Aging 2011, 3, 846–851. [Google Scholar] [CrossRef] [Green Version]
  384. Wu, Q.; Combs, C.; Cannady, S.B.; Geldmacher, D.S.; Herrup, K. Beta-amyloid activated microglia induce cell cycling and cell death in cultured cortical neurons. Neurobiol. Aging 2000, 21, 797–806. [Google Scholar] [CrossRef]
  385. Nagy, Z.; Esiri, M.M.; Smith, A.D. The cell division cycle and the pathophysiology of Alzheimer’s disease. Neuroscience 1998, 87, 731–739. [Google Scholar]
  386. Nagy, Z.; Esiri, M.M.; Cato, A.M.; Smith, A.D. Cell cycle markers in the hippocampus in Alzheimer’s disease. Acta Neuropathol. 1997, 94, 6–15. [Google Scholar] [CrossRef] [PubMed]
  387. Yang, Y.; Geldmacher, D.S.; Herrup, K. DNA replication precedes neuronal cell death in Alzheimer’s disease. J. Neurosci. 2001, 21, 2661–2668. [Google Scholar] [CrossRef] [PubMed]
  388. Yang, Y.; Mufson, E.J.; Herrup, K. Neuronal cell death is preceded by cell cycle events at all stages of Alzheimer’s disease. J. Neurosci. 2003, 23, 2557–2563. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  389. Tomashevski, A.; Husseman, J.; Jin, L.-W.; Nochlin, D.; Vincent, I. Constitutive Wee1 activity in adult brain neurons with M phase-type alterations in Alzheimer neurodegeneration. J. Alzheimers Dis. 2001, 3, 195–207. [Google Scholar] [CrossRef] [PubMed]
  390. Ding, X.L.; Husseman, J.; Tomashevski, A.; Nochlin, D.; Jin, L.W.; Vincent, I. The cell cycle Cdc25A tyrosine phosphatase is activated in degenerating postmitotic neurons in Alzheimer’s disease. Am. J. Pathol. 2000, 157, 1983–1990. [Google Scholar] [CrossRef]
  391. Milton, N.G.N. Phosphorylated amyloid-beta: The toxic intermediate in alzheimer’s disease neurodegeneration. Subcell Biochem. 2005, 38, 381–402. [Google Scholar]
  392. Kumar, S.; Wirths, O.; Stüber, K.; Wunderlich, P.; Koch, P.; Theil, S.; Rezaei-Ghaleh, N.; Zweckstetter, M.; Bayer, T.A.; Brüstle, O.; et al. Phosphorylation of the amyloid β-peptide at Ser26 stabilizes oligomeric assembly and increases neurotoxicity. Acta Neuropathol. 2016, 131, 525–537. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  393. Oumata, N.; Bettayeb, K.; Ferandin, Y.; Demange, L.; Lopez-Giral, A.; Goddard, M.L.; Myrianthopoulos, V.; Mikros, E.; Flajolet, M.; Greengard, P.; et al. Roscovitine-derived, dual-specificity inhibitors of cyclin-dependent kinases and casein kinases 1. J. Med Chem. 2008, 51, 5229–5242. [Google Scholar] [CrossRef] [PubMed]
  394. Baumann, K.; Mandelkow, E.M.; Biernat, J.; Piwnica-Worms, H.; Mandelkow, E. Abnormal Alzheimer-like phosphorylation of tau-protein by cyclin-dependent kinases cdk2 and cdk5. FEBS Lett. 1993, 336, 417–424. [Google Scholar] [CrossRef]
  395. Eichner, T.; Kutter, S.; Labeikovsky, W.; Buosi, V.; Kern, D. Molecular Mechanism of Pin1-Tau Recognition and Catalysis. J. Mol. Biol. 2016, 428 Pt A, 1760–1775. [Google Scholar] [CrossRef]
  396. Holzer, M.; Schade, N.; Opitz, A.; Hilbrich, I.; Stieler, J.; Vogel, T.; Neukel, V.; Oberstadt, M.; Totzke, F.; Schächtele, C.; et al. Novel Protein Kinase Inhibitors Related to Tau Pathology Modulate Tau Protein-Self Interaction Using a Luciferase Complementation Assay. Molecules 2018, 23, 2335. [Google Scholar] [CrossRef] [Green Version]
  397. Lee, K.H.; Lee, S.J.; Lee, H.J.; Choi, G.E.; Jung, Y.H.; Kim, D.I.; Gabr, A.A.; Ryu, J.M.; Han, H.J. Amyloid β1-42 (Aβ1-42) Induces the CDK2-Mediated Phosphorylation of Tau through the Activation of the mTORC1 Signaling Pathway While Promoting Neuronal Cell Death. Front. Mol. Neurosci. 2017, 10, 229. [Google Scholar] [CrossRef] [Green Version]
  398. Lee, M.S.; Kwon, Y.T.; Li, M.; Peng, J.; Friedlander, R.M.; Tsai, L.H. Neurotoxicity induces cleavage of p35 to p25 by calpain. Nature 2000, 405, 360–364. [Google Scholar] [CrossRef] [PubMed]
  399. Cruz, J.C.; Tsai, L.-H. Cdk5 deregulation in the pathogenesis of Alzheimer’s disease. Trends Mol. Med. 2004, 10, 452–458. [Google Scholar] [CrossRef]
  400. Cruz, J.C.; Tseng, H.-C.; Goldman, J.A.; Shih, H.; Tsai, L.-H. Aberrant Cdk5 activation by p25 triggers pathological events leading to neurodegeneration and neurofibrillary tangles. Neuron 2003, 40, 471–483. [Google Scholar] [CrossRef] [Green Version]
  401. Patrick, G.N.; Zukerberg, L.; Nikolic, M.; de la Monte, S.; Dikkes, P.; Tsai, L.H. Conversion of p35 to p25 deregulates Cdk5 activity and promotes neurodegeneration. Nature 1999, 402, 615–622. [Google Scholar] [CrossRef] [PubMed]
  402. Wang, Y.; Sheng, H.; Zhao, J.; Guo, L.; Liu, J.; Xu, J.; Liu, Q.; Huang, J.; Jiang, R.; Gan, S.; et al. Changes in the prefrontal cortex after the hippocampus was injected with Aβ25-35 via the P35/P25-CDK5-Tau hyperphosphorylation signaling pathway. Neurosci. Lett. 2021, 741, 135453. [Google Scholar] [CrossRef]
  403. Eftekharzadeh, B.; Daigle, J.G.; Kapinos, L.E.; Coyne, A.; Schiantarelli, J.; Carlomagno, Y.; Cook, C.; Miller, S.J.; Dujardin, S.; Amaral, A.S.; et al. Tau Protein Disrupts Nucleocytoplasmic Transport in Alzheimer’s Disease. Neuron 2018, 99, 925–940.e7. [Google Scholar] [CrossRef] [Green Version]
  404. Paonessa, F.; Evans, L.D.; Solanki, R.; Larrieu, D.; Wray, S.; Hardy, J.; Jackson, S.P.; Livesey, F.J. Microtubules Deform the Nuclear Membrane and Disrupt Nucleocytoplasmic Transport in Tau-Mediated Frontotemporal Dementia. Cell Rep. 2019, 26, 582–593.e5. [Google Scholar] [CrossRef] [Green Version]
  405. Giese, K.P. Generation of the Cdk5 activator p25 is a memory mechanism that is affected in early Alzheimer’s disease. Front. Mol. Neurosci. 2014, 7, 36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  406. Lau, L.-F.; Ahlijanian, M.K. Role of cdk5 in the pathogenesis of Alzheimer’s disease. Neurosignals 2003, 12, 209–214. [Google Scholar] [CrossRef] [PubMed]
  407. Liu, S.-L.; Wang, C.; Jiang, T.; Tan, L.; Xing, A.; Yu, J.-T. The Role of Cdk5 in Alzheimer’s Disease. Mol. Neurobiol. 2016, 53, 4328–4342. [Google Scholar] [CrossRef]
  408. Bhounsule, A.S.; Bhatt, L.K.; Prabhavalkar, K.S.; Oza, M. Cyclin dependent kinase 5: A novel avenue for Alzheimer’s disease. Brain Res. Bull. 2017, 132, 28–38. [Google Scholar] [CrossRef] [PubMed]
  409. Huang, Y.; Huang, W.; Huang, Y.; Song, P.; Zhang, M.; Zhang, H.T.; Pan, S.; Hu, Y. Cdk5 Inhibitory Peptide Prevents Loss of Neurons and Alleviates Behavioral Changes in p25 Transgenic Mice. J. Alzheimers Dis. 2020, 74, 1231–1242. [Google Scholar] [CrossRef] [PubMed]
  410. Sundaram, J.R.; Poore, C.P.; Sulaimee, N.H.B.; Pareek, T.; Asad, A.B.M.A.; Rajkumar, R.; Cheong, W.F.; Wenk, M.R.; Dawe, G.S.; Chuang, K.H.; et al. Specific inhibition of p25/Cdk5 activity by the Cdk5 inhibitory peptide reduces neurodegeneration in vivo. J. Neurosci. 2013, 33, 334–343. [Google Scholar] [CrossRef] [Green Version]
  411. He, Y.; Pan, S.; Xu, M.; He, R.; Huang, W.; Song, P.; Huang, J.; Zhang, H.T.; Hu, Y. Adeno-associated virus 9-mediated Cdk5 inhibitory peptide reverses pathologic changes and behavioral deficits in the Alzheimer’s disease mouse model. FASEB J. 2017, 31, 3383–3392. [Google Scholar] [CrossRef] [Green Version]
  412. Xu, M.; Huang, Y.; Song, P.; Huang, Y.; Huang, W.; Zhang, H.T.; Hu, Y. AAV9-Mediated Cdk5 Inhibitory Peptide Reduces Hyperphosphorylated Tau and Inflammation and Ameliorates Behavioral Changes Caused by Overexpression of p25 in the Brain. J. Alzheimers Dis. 2019, 70, 573–585. [Google Scholar] [CrossRef]
  413. Shukla, V.; Seo, J.; Binukumar, B.K.; Amin, N.D.; Reddy, P.; Grant, P.; Kuntz, S.; Kesavapany, S.; Steiner, J.; Mishra, S.K.; et al. TFP5, a Peptide Inhibitor of Aberrant and Hyperactive Cdk5/p25, Attenuates Pathological Phenotypes and Restores Synaptic Function in CK-p25Tg Mice. J. Alzheimers Dis. 2017, 56, 335–349. [Google Scholar] [CrossRef] [Green Version]
  414. Shukla, V.; Zheng, Y.L.; Mishra, S.K.; Amin, N.D.; Steiner, J.; Grant, P.; Kesavapany, S.; Pant, H.C. A truncated peptide from p35, a Cdk5 activator, prevents Alzheimer’s disease phenotypes in model mice. FASEB J. 2013, 27, 174–186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  415. Crews, L.; Patrick, C.; Adame, A.; Rockenstein, E.; Masliah, E. Modulation of aberrant CDK5 signaling rescues impaired neurogenesis in models of Alzheimer’s disease. Cell Death Dis. 2011, 2, e120. [Google Scholar] [CrossRef]
  416. Jorda, R.; Paruch, K.; Krystof, V. Cyclin-dependent kinase inhibitors inspired by roscovitine: Purine bioisosteres. Curr. Pharm. Des. 2012, 18, 2974–2980. [Google Scholar] [CrossRef]
  417. Chen, J.; Li, S.; Sun, W.; Li, J. Anti-diabetes drug pioglitazone ameliorates synaptic defects in AD transgenic mice by inhibiting cyclin-dependent kinase5 activity. PLoS ONE 2015, 10, e0123864. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  418. Shah, P.; Mudaliar, S. Pioglitazone: Side effect and safety profile. Expert Opin. Drug Saf. 2010, 9, 347–354. [Google Scholar] [CrossRef] [PubMed]
  419. Pelegrí, C.; Duran-Vilaregut, J.; del Valle, J.; Crespo-Biel, N.; Ferrer, I.; Pallàs, M.; Camins, A.; Vilaplana, J. Cell cycle activation in striatal neurons from Huntington’s disease patients and rats treated with 3-nitropropionic acid. Int. J. Dev. Neurosci. 2008, 26, 665–671. [Google Scholar] [CrossRef] [PubMed]
  420. Fernandez-Fernandez, M.R.; Ferrer, I.; Lucas, J.J. Impaired ATF6α processing, decreased Rheb and neuronal cell cycle re-entry in Huntington’s disease. Neurobiol. Dis. 2011, 41, 23–32. [Google Scholar] [CrossRef] [PubMed]
  421. Imray, F.P.; Kidson, C. Perturbations of cell-cycle progression in gamma-irradiated ataxia telangiectasia and Huntington’s disease cells detected by DNA flow cytometric analysis. Mutat. Res. 1983, 112, 369–382. [Google Scholar] [PubMed]
  422. Liu, K.Y.; Shyu, Y.C.; Barbaro, B.A.; Lin, Y.T.; Chern, Y.; Thompson, L.M.; James Shen, C.K.; Marsh, J.L. Disruption of the nuclear membrane by perinuclear inclusions of mutant huntingtin causes cell-cycle re-entry and striatal cell death in mouse and cell models of Huntington’s disease. Hum. Mol. Genet. 2015, 24, 1602–1616. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  423. Akashiba, H.; Ikegaya, Y.; Nishiyama, N.; Matsuki, N. Differential involvement of cell cycle reactivation between striatal and cortical neurons in cell death induced by 3-nitropropionic acid. J. Biol. Chem. 2008, 283, 6594–6606. [Google Scholar] [CrossRef] [Green Version]
  424. Chandrasekaran, S.; Bonchev, D. Network analysis of human post-mortem microarrays reveals novel genes, microRNAs, and mechanistic scenarios of potential importance in fighting huntington’s disease. Comput. Struct. Biotechnol. J. 2016, 14, 117–130. [Google Scholar] [CrossRef] [Green Version]
  425. Crespo-Biel, N.; Camins, A.; Pelegrí, C.; Vilaplana, J.; Pallàs, M.; Canudas, A.M. 3-Nitropropionic acid activates calpain/cdk5 pathway in rat striatum. Neurosci. Lett. 2007, 421, 77–81. [Google Scholar] [CrossRef]
  426. Duff, K.; Paulsen, J.; Mills, J.; Beglinger, L.J.; Moser, D.J.; Smith, M.M.; Langbehn, D.; Stout, J.; Queller, S.; Harrington, D.L. Mild cognitive impairment in prediagnosed Huntington disease. Neurology 2010, 75, 500–507. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  427. Lawrence, A.D.; Hodges, J.R.; Rosser, A.E.; Kershaw, A.; Ffrench-Constant, C.; Rubinsztein, D.C.; Robbins, T.W.; Sahakian, B.J. Evidence for specific cognitive deficits in preclinical Huntington’s disease. Brain 1998, 121 Pt 7, 329–341. [Google Scholar] [CrossRef] [Green Version]
  428. Lawrence, A.D.; Sahakian, B.J.; Hodges, J.R.; Rosser, A.E.; Lange, K.W.; Robbins, T.W. Executive and mnemonic functions in early Huntington’s disease. Brain 1996, 119 Pt 5, 1633–1645. [Google Scholar] [CrossRef] [Green Version]
  429. Alvarez-Periel, E.; Puigdellívol, M.; Brito, V.; Plattner, F.; Bibb, J.A.; Alberch, J.; Ginés, S. Cdk5 Contributes to Huntington’s Disease Learning and Memory Deficits via Modulation of Brain Region-Specific Substrates. Mol. Neurobiol. 2018, 55, 6250–6268. [Google Scholar] [CrossRef] [PubMed]
  430. Smith, M.M.; Mills, J.A.; Epping, E.A.; Westervelt, H.J.; Paulsen, J.S. PREDICT-HD Investigators of the Huntington Study Group. Depressive symptom severity is related to poorer cognitive performance in prodromal Huntington disease. Neuropsychology 2012, 26, 664–669. [Google Scholar] [CrossRef]
  431. Gray, M. Astrocytes in Huntington’s Disease. Adv. Exp. Med. Biol. 2019, 1175, 355–381. [Google Scholar]
  432. Park, K.H.J.; Lu, G.; Fan, J.; Raymond, L.A.; Leavitt, B.R. Decreasing Levels of the cdk5 Activators, p25 and p35, Reduces Excitotoxicity in Striatal Neurons. J. Huntingt. Dis. 2012, 1, 89–96. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  433. Reynolds, D.S.; Carter, R.J.; Morton, A.J. Dopamine modulates the susceptibility of striatal neurons to 3-nitropropionic acid in the rat model of Huntington’s disease. J. Neurosci. 1998, 18, 10116–10127. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  434. Luo, S.; Vacher, C.; Davies, J.E.; Rubinsztein, D.C. Cdk5 phosphorylation of huntingtin reduces its cleavage by caspases: Implications for mutant huntingtin toxicity. J. Cell Biol. 2005, 169, 647–656. [Google Scholar] [CrossRef]
  435. Anne, S.L.; Saudou, F.; Humbert, S. Phosphorylation of huntingtin by cyclin-dependent kinase 5 is induced by DNA damage and regulates wild-type and mutant huntingtin toxicity in neurons. J. Neurosci. 2007, 27, 7318–7328. [Google Scholar] [CrossRef] [Green Version]
  436. Chen, H.-M.; Wang, L.; D’Mello, S.R. A chemical compound commonly used to inhibit PKR, {8-(imidazol-4-ylmethylene)-6H-azolidino[5, 4-g] benzothiazol-7-one}, protects neurons by inhibiting cyclin-dependent kinase. Eur. J. Neurosci. 2008, 28, 2003–2016. [Google Scholar] [CrossRef] [Green Version]
  437. Damiano, M.; Galvan, L.; Déglon, N.; Brouillet, E. Mitochondria in Huntington’s disease. Biochim. Biophys. Acta 2010, 1802, 52–61. [Google Scholar] [CrossRef] [Green Version]
  438. Guo, X.; Disatnik, M.-H.; Monbureau, M.; Shamloo, M.; Mochly-Rosen, D.; Qi, X. Inhibition of mitochondrial fragmentation diminishes Huntington’s disease-associated neurodegeneration. J. Clin. Investig. 2013, 123, 5371–5388. [Google Scholar] [CrossRef] [PubMed]
  439. Park, K.H.J.; Franciosi, S.; Parrant, K.; Lu, G.; Leavitt, B.R. p35 hemizygosity activates Akt but does not improve motor function in the YAC128 mouse model of Huntington’s disease. Neuroscience 2017, 352, 79–87. [Google Scholar] [CrossRef] [PubMed]
  440. Kaminosono, S.; Saito, T.; Oyama, F.; Ohshima, T.; Asada, A.; Nagai, Y.; Nukina, N.; Hisanaga, S.I. Suppression of mutant Huntingtin aggregate formation by Cdk5/p35 through the effect on microtubule stability. J. Neurosci. 2008, 28, 8747–8755. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  441. Mathys, H.; Davila-Velderrain, J.; Peng, Z.; Gao, F.; Mohammadi, S.; Young, J.Z.; Menon, M.; He, L.; Abdurrob, F.; Jiang, X.; et al. Single-cell transcriptomic analysis of Alzheimer’s disease. Nature 2019, 570, 332–337. [Google Scholar] [CrossRef]
  442. Zhou, Y.; Song, W.M.; Andhey, P.S.; Swain, A.; Levy, T.; Miller, K.R.; Poliani, P.L.; Cominelli, M.; Grover, S.; Gilfillan, S.; et al. Human and mouse single-nucleus transcriptomics reveal TREM2-dependent and TREM2-independent cellular responses in Alzheimer’s disease. Nat. Med. 2020, 26, 131–142. [Google Scholar] [CrossRef]
  443. Murray, M.E.; Graff-Radford, N.R.; Ross, O.A.; Petersen, R.C.; Duara, R.; Dickson, D.W. Neuropathologically defined subtypes of Alzheimer’s disease with distinct clinical characteristics: A retrospective study. Lancet Neurol. 2011, 10, 785–796. [Google Scholar] [CrossRef] [Green Version]
  444. Vogel, J.W.; Young, A.L.; Oxtoby, N.P.; Smith, R.; Ossenkoppele, R.; Strandberg, O.T.; La Joie, R.; Aksman, L.M.; Grothe, M.J.; Iturria-Medina, Y.; et al. Four distinct trajectories of tau deposition identified in Alzheimer’s disease. Nat. Med. 2021, 27, 1–11. [Google Scholar] [CrossRef]
  445. Ferrer, I. Oligodendrogliopathy in neurodegenerative diseases with abnormal protein aggregates: The forgotten partner. Prog. Neurobiol. 2018, 169, 24–54. [Google Scholar] [CrossRef] [Green Version]
  446. Kitabayashi, T.; Dong, Y.; Furuta, T.; Sabit, H.; Jiapaer, S.; Zhang, J.; Zhang, G.; Hayashi, Y.; Kobayashi, M.; Domoto, T.; et al. Identification of GSK3β inhibitor kenpaullone as a temozolomide enhancer against glioblastoma. Sci. Rep. 2019, 9, 10049. [Google Scholar] [CrossRef]
  447. Myre, M.A.; Washicosky, K.; Moir, R.D.; Tesco, G.; Tanzi, R.E.; Wasco, W. Reduced amyloidogenic processing of the amyloid beta-protein precursor by the small-molecule Differentiation Inducing Factor-1. Cell Signal. 2009, 21, 567–576. [Google Scholar] [CrossRef] [Green Version]
  448. Ma, X.H.; Wang, R.; Tan, C.Y.; Jiang, Y.Y.; Lu, T.; Rao, H.B.; Li, X.Y.; Go, M.L.; Low, B.C.; Chen, Y.Z. Virtual screening of selective multitarget kinase inhibitors by combinatorial support vector machines. Mol. Pharm. 2010, 7, 1545–1560. [Google Scholar] [CrossRef]
  449. Logé, C.; Testard, A.; Thiéry, V.; Lozach, O.; Blairvacq, M.; Robert, J.M.; Meijer, L.; Besson, T. Novel 9-oxo-thiazolo[5, 4-f]quinazoline-2-carbonitrile derivatives as dual cyclin-dependent kinase 1 (CDK1)/glycogen synthase kinase-3 (GSK-3) inhibitors: Synthesis, biological evaluation and molecular modeling studies. Eur. J. Med. Chem. 2008, 43, 1469–1477. [Google Scholar] [CrossRef]
  450. Mayes, P.A.; Dolloff, N.G.; Daniel, C.J.; Liu, J.J.; Hart, L.S.; Kuribayashi, K.; Allen, J.E.; Jee, D.I.; Dorsey, J.F.; Liu, Y.Y.; et al. Overcoming hypoxia-induced apoptotic resistance through combinatorial inhibition of GSK-3β and CDK1. Cancer Res. 2011, 71, 5265–5275. [Google Scholar] [CrossRef] [Green Version]
  451. Kunick, C.; Lauenroth, K.; Wieking, K.; Xie, X.; Schultz, C.; Gussio, R.; Zaharevitz, D.; Leost, M.; Meijer, L.; Weber, A.; et al. Evaluation and comparison of 3D-QSAR CoMSIA models for CDK1, CDK5, and GSK-3 inhibition by paullones. J. Med. Chem. 2004, 47, 22–36. [Google Scholar] [CrossRef] [PubMed]
  452. Tell, V.; Holzer, M.; Herrmann, L.; Mahmoud, K.A.; Schächtele, C.; Totzke, F.; Hilgeroth, A. Multitargeted drug development: Discovery and profiling of dihydroxy substituted 1-aza-9-oxafluorenes as lead compounds targeting Alzheimer disease relevant kinases. Bioorg. Med. Chem. Lett. 2012, 22, 6914–6918. [Google Scholar] [CrossRef] [PubMed]
  453. Ortega, M.A.; Montoya, M.E.; Zarranz, B.; Jaso, A.; Aldana, I.; Leclerc, S.; Meijer, L.; Monge, A. Pyrazolo[3, 4-b]quinoxalines. A new class of cyclin-dependent kinases inhibitors. Bioorg. Med. Chem. 2002, 10, 2177–2184. [Google Scholar] [CrossRef]
  454. Mettey, Y.; Gompel, M.; Thomas, V.; Garnier, M.; Leost, M.; Ceballos-Picot, I.; Noble, M.; Endicott, J.; Vierfond, J.M.; Meijer, L. Aloisines, a new family of CDK/GSK-3 inhibitors. SAR study, crystal structure in complex with CDK2, enzyme selectivity, and cellular effects. J. Med. Chem. 2003, 46, 222–236. [Google Scholar] [CrossRef]
  455. Heider, F.; Ansideri, F.; Tesch, R.; Pantsar, T.; Haun, U.; Döring, E.; Kudolo, M.; Poso, A.; Albrecht, W.; Laufer, S.A.; et al. Pyridinylimidazoles as dual glycogen synthase kinase 3β/p38α mitogen-activated protein kinase inhibitors. Eur. J. Med. Chem. 2019, 175, 309–329. [Google Scholar] [CrossRef]
  456. Ouach, A.; Boulahjar, R.; Vala, C.; Bourg, S.; Bonnet, P.; Guguen-Guillouzo, C.; Ravache, M.; Le Guevel, R.; Lozach, O.; Lazar, S.; et al. Novel optimization of valmerins (tetrahydropyrido[1,8 2-a]isoindolones) as potent dual CDK5/GSK3 inhibitors. Eur. J. Med. Chem. 2016, 115, 311–325. [Google Scholar] [CrossRef]
  457. Li, X.; Wang, X.; Tian, Z.; Zhao, H.; Liang, D.; Li, W.; Qiu, Y.; Lu, S. Structural basis of valmerins as dual inhibitors of GSK3β/CDK5. J. Mol. Model. 2014, 20, 2407. [Google Scholar] [CrossRef]
  458. Reinhardt, L.; Kordes, S.; Reinhardt, P.; Glatza, M.; Baumann, M.; Drexler, H.C.; Menninger, S.; Zischinsky, G.; Eickhoff, J.; Fröb, C.; et al. Dual Inhibition of GSK3β and CDK5 Protects the Cytoskeleton of Neurons from Neuroinflammatory-Mediated Degeneration In Vitro and In Vivo. Stem Cell Rep. 2019, 12, 502–517. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  459. Wang, L.; Ankati, H.; Akubathini, S.K.; Balderamos, M.; Storey, C.A.; Patel, A.V.; Price, V.; Kretzschmar, D.; Biehl, E.R.; D’Mello, S.R. Identification of novel 1, 4-benzoxazine compounds that are protective in tissue culture and in vivo models of neurodegeneration. J. Neurosci. Res. 2010, 88, 1970–1984. [Google Scholar] [CrossRef] [PubMed]
  460. Liu, J.; Yang, J.; Xu, Y.; Guo, G.; Cai, L.; Wu, H.; Zhao, Y.; Zhang, X. Roscovitine, a CDK5 Inhibitor, Alleviates Sevoflurane-Induced Cognitive Dysfunction via Regulation Tau/GSK3β and ERK/PPARγ/CREB Signaling. Cell Physiol. Biochem. 2017, 44, 423–435. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  461. Chow, H.M.; Guo, D.; Zhou, J.C.; Zhang, G.Y.; Li, H.F.; Herrup, K.; Zhang, J. CDK5 activator protein p25 preferentially binds and activates GSK3β. Proc. Natl. Acad. Sci. USA 2014, 111, E4887–E4895. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  462. Kockeritz, L.; Doble, B.; Patel, S.; Woodgett, J.R. Glycogen synthase kinase-3-an overview of an over-achieving protein kinase. Curr. Drug Targets. 2006, 7, 1377–1388. [Google Scholar] [CrossRef]
  463. Cohen, Y.; Chetrit, A.; Cohen, Y.; Sirota, P.; Modan, B. Cancer morbidity in psychiatric patients: Influence of lithium carbonate treatment. Med. Oncol. 1998, 15, 32–36. [Google Scholar] [CrossRef]
  464. Salman, M.M.; Al-Obaidi, Z.; Kitchen, P.; Loreto, A.; Bill, R.M.; Wade-Martins, R. Advances in Applying Computer-Aided Drug Design for Neurodegenerative Diseases. Int. J. Mol. Sci. 2021, 22, 4688. [Google Scholar] [CrossRef]
  465. Makhouri, F.R.; Ghasemi, J.B. In Silico Studies in Drug Research Against Neurodegenerative Diseases. Curr. Neuropharmacol. 2018, 16, 664–725. [Google Scholar] [CrossRef]
  466. Carpenter, K.A.; Huang, X. Machine Learning-based Virtual Screening and Its Applications to Alzheimer’s Drug Discovery: A Review. Curr. Pharm. Des. 2018, 24, 3347–3358. [Google Scholar] [CrossRef]
  467. Aldewachi, H.; Al-Zidan, R.N.; Conner, M.T.; Salman, M.M. High-Throughput Screening Platforms in the Discovery of Novel Drugs for Neurodegenerative Diseases. Bioengineering 2021, 8, 30. [Google Scholar] [CrossRef] [PubMed]
  468. Varma, H.; Lo, D.C.; Stockwell, B.R. High-Throughput and High-Content Screening for Huntington’s Disease Therapeutics. In Neurobiology of Huntington’s Disease: Applications to Drug Discovery; Lo, D.C., Hughes, R.E., Eds.; CRC Press/Taylor & Francis: Boca Raton, FL, USA, 2011. Available online: http://www.ncbi.nlm.nih.gov/books/NBK55989/ (accessed on 20 May 2021).
  469. Lublin, A.L.; Link, C.D. Alzheimer’s disease drug discovery: In vivo screening using Caenorhabditis elegans as a model for β-amyloid peptide-induced toxicity. Drug Discov. Today Technol. 2013, 10, e115–e119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  470. Lee, C.-T.; Bendriem, R.M.; Wu, W.W.; Shen, R.-F. 3D brain Organoids derived from pluripotent stem cells: Promising experimental models for brain development and neurodegenerative disorders. J. Biomed. Sci. 2017, 24, 59. [Google Scholar] [CrossRef] [PubMed]
  471. Choi, S.H.; Kim, Y.H.; Quinti, L.; Tanzi, R.E.; Kim, D.Y. 3D culture models of Alzheimer’s disease: A road map to a “cure-in-a-dish.”. Mol. Neurodegener. 2016, 11, 75. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  472. Kumar, A.; Zhou, L.; Zhi, K.; Raji, B.; Pernell, S.; Tadrous, E.; Kodidela, S.; Nookala, A.; Kochat, H.; Kumar, S. Challenges in Biomaterial-Based Drug Delivery Approach for the Treatment of Neurodegenerative Diseases: Opportunities for Extracellular Vesicles. Int. J. Mol. Sci. 2020, 22, 138. [Google Scholar] [CrossRef] [PubMed]
  473. Grabrucker, A.M.; Ruozi, B.; Belletti, D.; Pederzoli, F.; Forni, F.; Vandelli, M.A.; Tosi, G. Nanoparticle transport across the blood brain barrier. Tissue Barriers 2016, 4, e1153568. [Google Scholar] [CrossRef] [Green Version]
  474. Karthivashan, G.; Ganesan, P.; Park, S.-Y.; Kim, J.-S.; Choi, D.-K. Therapeutic strategies and nano-drug delivery applications in management of ageing Alzheimer’s disease. Drug Deliv. 2018, 25, 307–320. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  475. Lee, S.W.L.; Paoletti, C.; Campisi, M.; Osaki, T.; Adriani, G.; Kamm, R.D.; Mattu, C.; Chiono, V. MicroRNA delivery through nanoparticles. J. Control Release 2019, 313, 80–95. [Google Scholar] [CrossRef]
  476. Akel, H.; Ismail, R.; Csóka, I. Progress and perspectives of brain-targeting lipid-based nanosystems via the nasal route in Alzheimer’s disease. Eur. J. Pharm. Biopharm. 2020, 148, 38–53. [Google Scholar] [CrossRef] [PubMed]
  477. Battaglia, L.; Panciani, P.P.; Muntoni, E.; Capucchio, M.T.; Biasibetti, E.; De Bonis, P.; Mioletti, S.; Fontanella, M.; Swaminathan, S. Lipid nanoparticles for intranasal administration: Application to nose-to-brain delivery. Expert Opin. Drug Deliv. 2018, 15, 369–378. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Effects of elevated GSK3, p38 MAPK and CDK5 activity in models of AD. (A) GSK3 increases Aβ levels by activating PS1 [165,166,167] and increasing the expression of BACE1 [168,169], impairs Aβ and Tau clearance by inhibiting autophagy [172,173], promotes neuroinflammation [147,149,151,174], depresses LTP [175], and negatively regulates synaptic plasticity and learning/memory [176,177]. Through E2F1 phosphorylation, GSK3 promotes abortive cell cycle entry [178,179] and by phosphorylation Tau causes its disassociation from microtubules and assembly into fibrils and NFTs [15,158,159,160,161]. GSK3 also inhibits adult neurogenesis and promotes neuronal death in the hippocampus [143]. Although inhibiting GSK3 protects in experimental models of HD, less is known about the mechanism by which elevated GSK3 activity promotes neuronal death in HD. Several lines of evidence indicate that GSK3 activates HDAC3-mediated neurotoxicity through its phosphorylation [106]. GSK3 also activates caspase-3, phosphorylates Tau, promotes neuroinflammation and causes cognitive impairment in HD models [180]. (B) In astrocytes and microglia p38 MAP promotes release of inflammatory cytokines resulting in chronic neuroinflammation and consequently to neurodegeneration [181,182,183]. Within neurons p38 has been shown to have several effects that promote neurodegeneration including the activation of apoptotic signaling pathways [182,184,185,186], the phosphorylation of Tau [187,188,189], the phosphorylation of E2F1 [190,191], inhibition of LTP [192,193,194], the promotion of excitotoxicity through activation of the NMDA glutamate receptor [195,196,197], and the inhibition of autophagy through the phosphorylation of the ULK1 complex [198,199,200], which impairs the ability of neurons and glial cells to clear Tau and Aβ aggregates. It should be noted that the actions depicted in the figure pertain to p38α, p38βand p38δ MAPK, and most specifically to p38α. In contrast to the other p38 MAPKs, p38γ is believed to have neuroprotective actions in the context of AD. Although inhibiting p38 MAPK promotes neurodegeneration in HD by acting both in neurons and in glial cells, the mechanisms involved are unclear. Most emphasis has been placed on the mechanisms by which p38 MAPK activity is increased in HD. (C) Not shown in the figure are the actions of cell cycle-promoting CDKs, which are aberrantly activated in neurons in the AD brain resulting in an abortive entry into the cell cycle culminating in cell death [34,35,36,201,202,203]. Cell cycle promoting CDKs also promote neurodegeneration by other mechanisms, such as increasing APP processing and inhibiting autophagy [204]. Of the CDKs, CDK5 is most involved on AD. By stimulating PS1 and BACE1 activity, CDK5 increases Aβ levels by both transcriptional mechanisms and through PS1 and BACE1 activation [205,206,207]. Elevated Aβ, through activation of CDK1, 2 and 4, leads to the phosphorylation of lamins and nuclear membrane damage [208]. CDK5 is also a major Tau kinase and causes Tau dysfunction [209]. Through phosphorylation of Vps35, CDK inhibits autophagic clearance of Aβ and Tau [204,210,211]. CDK5 produces neuroinflammation both by promoting neuronal death and more directly by stimulating lysophosphatidylcholine release from glia [212]. CDK5 also plays a key role in HD pathogenesis by inducing mitochondrial fission [213], generating oxidative stress and promoting excitotoxicity [214,215], by the aberrant phosphorylation of DARPP32 [216], by increasing expression of the pro-apoptotic protein, c-jun [217], and increasing activity of specific cell cycle-promoting CDKs through the stimulation of expression of their cognate cyclins [218].
Figure 1. Effects of elevated GSK3, p38 MAPK and CDK5 activity in models of AD. (A) GSK3 increases Aβ levels by activating PS1 [165,166,167] and increasing the expression of BACE1 [168,169], impairs Aβ and Tau clearance by inhibiting autophagy [172,173], promotes neuroinflammation [147,149,151,174], depresses LTP [175], and negatively regulates synaptic plasticity and learning/memory [176,177]. Through E2F1 phosphorylation, GSK3 promotes abortive cell cycle entry [178,179] and by phosphorylation Tau causes its disassociation from microtubules and assembly into fibrils and NFTs [15,158,159,160,161]. GSK3 also inhibits adult neurogenesis and promotes neuronal death in the hippocampus [143]. Although inhibiting GSK3 protects in experimental models of HD, less is known about the mechanism by which elevated GSK3 activity promotes neuronal death in HD. Several lines of evidence indicate that GSK3 activates HDAC3-mediated neurotoxicity through its phosphorylation [106]. GSK3 also activates caspase-3, phosphorylates Tau, promotes neuroinflammation and causes cognitive impairment in HD models [180]. (B) In astrocytes and microglia p38 MAP promotes release of inflammatory cytokines resulting in chronic neuroinflammation and consequently to neurodegeneration [181,182,183]. Within neurons p38 has been shown to have several effects that promote neurodegeneration including the activation of apoptotic signaling pathways [182,184,185,186], the phosphorylation of Tau [187,188,189], the phosphorylation of E2F1 [190,191], inhibition of LTP [192,193,194], the promotion of excitotoxicity through activation of the NMDA glutamate receptor [195,196,197], and the inhibition of autophagy through the phosphorylation of the ULK1 complex [198,199,200], which impairs the ability of neurons and glial cells to clear Tau and Aβ aggregates. It should be noted that the actions depicted in the figure pertain to p38α, p38βand p38δ MAPK, and most specifically to p38α. In contrast to the other p38 MAPKs, p38γ is believed to have neuroprotective actions in the context of AD. Although inhibiting p38 MAPK promotes neurodegeneration in HD by acting both in neurons and in glial cells, the mechanisms involved are unclear. Most emphasis has been placed on the mechanisms by which p38 MAPK activity is increased in HD. (C) Not shown in the figure are the actions of cell cycle-promoting CDKs, which are aberrantly activated in neurons in the AD brain resulting in an abortive entry into the cell cycle culminating in cell death [34,35,36,201,202,203]. Cell cycle promoting CDKs also promote neurodegeneration by other mechanisms, such as increasing APP processing and inhibiting autophagy [204]. Of the CDKs, CDK5 is most involved on AD. By stimulating PS1 and BACE1 activity, CDK5 increases Aβ levels by both transcriptional mechanisms and through PS1 and BACE1 activation [205,206,207]. Elevated Aβ, through activation of CDK1, 2 and 4, leads to the phosphorylation of lamins and nuclear membrane damage [208]. CDK5 is also a major Tau kinase and causes Tau dysfunction [209]. Through phosphorylation of Vps35, CDK inhibits autophagic clearance of Aβ and Tau [204,210,211]. CDK5 produces neuroinflammation both by promoting neuronal death and more directly by stimulating lysophosphatidylcholine release from glia [212]. CDK5 also plays a key role in HD pathogenesis by inducing mitochondrial fission [213], generating oxidative stress and promoting excitotoxicity [214,215], by the aberrant phosphorylation of DARPP32 [216], by increasing expression of the pro-apoptotic protein, c-jun [217], and increasing activity of specific cell cycle-promoting CDKs through the stimulation of expression of their cognate cyclins [218].
Ijms 22 05911 g001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

D’Mello, S.R. When Good Kinases Go Rogue: GSK3, p38 MAPK and CDKs as Therapeutic Targets for Alzheimer’s and Huntington’s Disease. Int. J. Mol. Sci. 2021, 22, 5911. https://doi.org/10.3390/ijms22115911

AMA Style

D’Mello SR. When Good Kinases Go Rogue: GSK3, p38 MAPK and CDKs as Therapeutic Targets for Alzheimer’s and Huntington’s Disease. International Journal of Molecular Sciences. 2021; 22(11):5911. https://doi.org/10.3390/ijms22115911

Chicago/Turabian Style

D’Mello, Santosh R. 2021. "When Good Kinases Go Rogue: GSK3, p38 MAPK and CDKs as Therapeutic Targets for Alzheimer’s and Huntington’s Disease" International Journal of Molecular Sciences 22, no. 11: 5911. https://doi.org/10.3390/ijms22115911

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