Next Article in Journal
Loss of Human Beta Cell Identity in a Reconstructed Omental Stromal Cell Environment
Next Article in Special Issue
Populations of Tau Conformers Drive Prion-like Strain Effects in Alzheimer’s Disease and Related Dementias
Previous Article in Journal
Biomarkers Associated with Cardiovascular Disease in COVID-19
Previous Article in Special Issue
Impairments in Brain Bioenergetics in Aging and Tau Pathology: A Chicken and Egg Situation?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Microtubules as Regulators of Neural Network Shape and Function: Focus on Excitability, Plasticity and Memory

by
Fernando Peña-Ortega
*,
Ángel Abdiel Robles-Gómez
and
Lorena Xolalpa-Cueva
Departamento de Neurobiología del Desarrollo y Neurofisiología, Instituto de Neurobiología, Universidad Nacional Autónoma de México, Campus Juriquilla, Boulevard Juriquilla 3001, Querétaro 76230, Mexico
*
Author to whom correspondence should be addressed.
Cells 2022, 11(6), 923; https://doi.org/10.3390/cells11060923
Submission received: 1 January 2022 / Revised: 9 February 2022 / Accepted: 17 February 2022 / Published: 8 March 2022
(This article belongs to the Collection Tau Protein in Neurodegenerative Diseases)

Abstract

:
Neuronal microtubules (MTs) are complex cytoskeletal protein arrays that undergo activity-dependent changes in their structure and function as a response to physiological demands throughout the lifespan of neurons. Many factors shape the allostatic dynamics of MTs and tubulin dimers in the cytosolic microenvironment, such as protein–protein interactions and activity-dependent shifts in these interactions that are responsible for their plastic capabilities. Recently, several findings have reinforced the role of MTs in behavioral and cognitive processes in normal and pathological conditions. In this review, we summarize the bidirectional relationships between MTs dynamics, neuronal processes, and brain and behavioral states. The outcomes of manipulating the dynamicity of MTs by genetic or pharmacological approaches on neuronal morphology, intrinsic and synaptic excitability, the state of the network, and behaviors are heterogeneous. We discuss the critical position of MTs as responders and adaptative elements of basic neuronal function whose impact on brain function is not fully understood, and we highlight the dilemma of artificially modulating MT dynamics for therapeutic purposes.

Graphical Abstract

1. Introduction

Neurons are postmitotic, highly polarized cells with complex morphological and functional compartments, such as soma, axon, dendrites, and synapses [1] (Figure 1). These compartments are supported morphologically and functionally on specialized and specific cytoskeletal arrangements [2,3,4,5] (Figure 1). The cytoskeleton comprises filamentous actin, intermediate filaments, microtubules (MTs), and associated proteins [6,7]. MTs are particularly important because they support the neuronal complex and dynamic branching and compartmentalization while acting as an intracellular roadmap for protein motors to deliver important cargoes (e.g., receptors, neurotransmitters) and organelles to various cell regions [1] and elements that modulate (and are modulated by) fundamental neural functions [1,8], as will be discussed in this review.
MTs are dynamic, cylindrical, linear, and anionic polymers with a hollow tubular structure. They have variable lengths [9] and are composed of α- and β-tubulin, which interact noncovalently to form a stable heterodimer with polarity orientation [10,11], a fast-growing plus end, and a slow-growing minus end [12]. MT assembly requires α- and β-tubulin to be bound to GTP, which is permanently bound to α-tubulin [13,14] and hydrolyzed to GDP by β-tubulin after tubulin dimers assembly into the MT [15,16]. The GDP attached to β-tubulin incorporated into MTs can only be replaced by GTP when tubulin returns to its heterodimeric form during the continuous cycles of polymerization and depolymerization [9].
MTs in differentiated neurons are not attached to the centrosome [17,18], which allows them to exist in different neuronal compartments as autonomous structures [19,20]. In neurons, most axonal MTs have their plus ends directed away from the cell body [21]. Their cargo transport is controlled in a directional manner and has kinesin motors carrying cargo toward axon terminals (anterograde) and dynein motors carrying cargo away from axon tips (retrograde) [22]. As will be reviewed later, MT localization at the presynaptic terminals has been recently accepted [23,24] (Figure 1). In contrast to axons, dendrites have a mixed MT orientation with both minus and plus ends oriented toward the cell body [20,25]. Although a fraction of the axonal MT array is indeed quite stable, a fraction close to the terminal is not, with dendritic MTs having a lower fraction of stable MT polymers than axons [26,27].
Neurons possess stable and dynamic pools of MTs that exhibit complex behavior, with some MTs in a growing phase, others stationary, and some in a state of disassembly [7]. The stochastic switch between growth to shortening (named catastrophe) and shortening to growth (named rescue) precludes MTs from reaching a steady-state length and makes them a structure that simultaneously undergoes assembly and disassembly, producing a condition named “dynamic instability” [28,29,30]. Dynamic instability is controlled by GTP hydrolysis and, thus, is an energy-consuming process [31] that varies depending on the isoforms of α- and β-tubulin incorporated into the MTs, their posttranslational modifications, and their interaction with MT associated proteins (MAPs) [32,33]. Despite being constantly changing structures, MTs have sufficient longevity to be substrates for tubulin modifying enzymes that lead to their tyrosination, detyrosination, acetylation, D2 modification, glutamylation, glycation, palmitoylation, and phosphorylation [34,35,36,37,38]. These posttranslational modifications modulate their binding to particular MAPs, motor proteins, or proteases [39,40,41,42], which are subject to these same posttranslational regulations.
MAPs constitute a class of heterogeneous regulators that modulate MT stability and dynamics, the directional transport of cargo, and MT nucleation. They also dynamically interact with other cellular proteins and organelles [43,44]. The brain expresses several MAPs such as Tau, MAP1A, MAP1B, MAP2, and MAP6, with some exhibiting a compartment-specific distribution. For example, MAP2A and MAP2B are preferentially located in the cell bodies and dendrites of mature neurons [45], whereas Tau protein and MAP6 are distributed mostly in the axonal compartment [46] and, to a lesser extent, in dendrites [47,48]. MAPs play a major role in the induction of distinctive morphologies between axons and dendrites [49,50], but they also regulate axonal transport [51] and neuronal plasticity [52] (Figure 1).
As will be reviewed here, MTs are fundamental in a variety of neural functions including neuronal excitability, synaptic coupling, synaptic plasticity, and memory, which are strongly dependent on a constant shift between MT stability and instability [53,54,55,56,57,58,59]. Thus, neurons are particularly susceptible to MT deregulation and defects closely related to neurological disorders [60,61,62]. MT destabilization has been associated to amyotrophic lateral sclerosis [63,64], Parkinson’s disease [65], Huntington’s disease [66], frontotemporal dementia with parkinsonism linked to chromosome 17 [67], Alzheimer’s disease (AD) [68] or psychiatric diseases [34,69,70,71,72] (Figure 1). Moreover, the relevance of MTs in brain function has also been reflected in the neuropathy induced by several MT stabilizers when used as anticancer drugs [73,74]. All these phenomena indicate that there is a delicate balance in MT stability/instability that is required for proper brain function [57]. Thus, the purpose of this review is to reveal the role of MTs and their endogenous and exogenous stability modulators in neural function and to shape the impact of neural function on MT configuration.

2. Modulation of MT Stability and Its Impact on Brain Function

Fibroblast growth factor 13 (FGF13), is an endogenous MT stabilizer. It is a non-secretory protein of the FGF family that acts intracellularly as an MT-stabilizing protein by promoting tubulin polymerization; its deletion leads to alterations in learning and memory [75]. As mentioned, MAPs bind and stabilize MTs in a phosphorylation-dependent manner [45,46,76] and their alterations disrupt brain function. For instance, the deletion of MAP6 triggers various neurotransmission and behavioral defects, leading to schizoaffective disorder, which could be corrected by the pharmacological stabilization of MTs [77]. The most prominent MAP is tau, which is encoded by the MAPT (microtubule-associated protein tau) gene [78], which generates six isoforms through alternative splicing [79]. Tau binds tubulin in a phosphorylation-dependent manner via its MT-binding domains [80], with a single Tau molecule crosslinking multiple tubulin dimers [81], which stabilizes MTs [82]. Although Tau is preferentially expressed within the axon, increasing its concentration towards the distal end [27], it is also located in the soma and dendrites [4,5,49,83,84]. Phosphorylation of tau, which is tightly regulated under physiological conditions [85,86], reduces its affinity for MTs, promoting their dismantling [87]. Tau hyperphosphorylation leads to aggregation and MT disruption [88,89,90,91], which is mainly due to its reduced MT-stabilizing properties but also to the sequestration of other MAPs [69,91], ultimately inducing neuronal dysfunction [92,93]. However, as will be reviewed later, blocking Tau phosphorylation also leads to neuronal morphological and functional alteration [5,94], which indicates that Tau phosphorylation has specific physiological functions, such as the modulation of N-methyl-D-aspartate (NMDA)-mediated processes [5,84,95]. Research on the role of MTs and associated proteins in brain function has paved the way for the use of pharmacological agents that modulate their stability. These agents have revealed the functional consequences of changes in MT stability under normal and pathological conditions.

3. Pharmacological Modulators of MT Stability

Microtubule stabilizing agents (MSAs), are among the most clinically used chemotherapeutic drugs [96] because they can inhibit cell division by stabilizing MTs [97,98]. However, MSAs have also been considered potential candidates for the treatment of neurological alterations related to MT destabilization [99,100,101,102]. Many MSAs of natural origin, including taxanes and epothilones (and their derivatives), have been approved for cancer treatment, but their use for neurological alterations has been halted due to their limited brain penetration, poor bioavailability, and/or their potential systemic side effects [103,104,105,106]. Most of these drugs interact with MTs at the taxane-binding site, located in the lumen of the MT in the β-tubulin subunit [107,108,109], counteracting the effects of its GTPase activity [110,111], protecting against MT depolarization and dissolution, and promoting the polarization and structural stability of MTs [112,113,114,115].
Paclitaxel (PTX) is one taxane used in chemotherapy [116,117], that is highly lipophilic and easily crosses the blood–brain barrier (BBB), but it is quickly eliminated from the CNS by p-glycoprotein-mediated transport [118,119,120]. Despite its poor CNS bioavailability, several studies have indicated that PTX could regulate neural shape and function (Figure 2). For instance, PTX promotes axonal elongation/regeneration and reduces glial scar formation in animal models of nerve injury [113,114]. PTX restores axonal transport and reduces the motor phenotype in transgenic mice exhibiting hyperphosphorylated Tau [90], which was linked to MT stabilization because of the increased levels of detyrosinated tubulin [90]. PTX also reduces glutamate-induced neurotoxicity [121] (Figure 2). Despite all these neuroprotective effects, PTX also induces peripheral neuropathy in humans, which is characterized by sensory abnormalities [122], including pain [123] and chemotherapy-induced cognitive impairment [124,125], which has been reproduced in animal models [126] and can also be induced with other taxane-related chemotherapeutic agents such as docetaxel (DTX) [127,128,129] (Figure 2). In rats, PTX treatment induces allodynia that correlates with altered brain activity and connectivity [126], which might be related to the PTX-induced transient encephalopathy documented in humans [130]. Moreover, PTX can induce neuronal death accompanied by the loss of MAP2 and the presence of dystrophic neurites [131] (Figure 2).
Epothilone D (Epo-D), also called BMS-241027, KOS862, dEpoB, and CRND66, is a PTX-derived brain-penetrant MT stabilizer [97], also used in chemotherapy [132,133,134], that prevents MT disassembly by interacting with β-tubulin at the taxane-binding site. This compound is promising to treat neurological diseases for its excellent BBB penetration [112,135] and its poor transport by the p-glycoprotein [97,132]. Epo-D can promote axonal sprouting in injured cortical neurons [136] and facilitate recovery of hind limb function after spinal cord injury in rats [137]. Epo-D improves outcomes in the MPTP-induced mouse model of parkinsonism [138] and in transgenic models of AD [48,111,112,139], by increasing MT density, axonal integrity, and neuronal survival. Epo-D reverts the behavioral alterations in MAP6 knockout mice [77], which correlated with an increase in synapse density and improved synaptic long-term potentiation (LTP) [135]. Moreover, Epo-D attenuates Tau pathology, improves MT density, attenuates axonal dystrophy, improves axonal transport, and enhances cognition in a transgenic mouse model of tauopathy [112,139]. Epo-D also restores normal MT dynamics in conditions of Tau disruption [54,139]. Moreover, Epo-D restores the density of mushroom spines affected by lateral fluid percussion brain injury [140]. Some of these neuroprotective effects have been reproduced with epothilone B (Epo-B) [138,141,142,143]. A collection of studies also report neurological adverse effects of epothilones in animal models [94,140,144]. For instance, Epo-D reduces dendritic arborization [94] and the viability of neurons by affecting mitochondrial transport [145], accelerating disease progression in a transgenic model of ALS [145]. Similar alterations can be induced with Epo-B in cortical and adult sensory neurons [142]. Interestingly, epothilone’s efficacy appears to be more effective in younger rather than aged animals with traumatic brain injury by preventing axonal degeneration processes [140,141,142,143,144,145,146,147,148]. Moreover, epothilones produce more neurotoxic effects in aged animals [140,148]. Thus, specific factors like age can determine the net physiological outcomes of some MT stabilizers. Sex differences in axon diameter, MT density, and resistance to stretch injury [148] could also bias the beneficial effects of MT stabilizers. However, this possibility has not yet been properly assessed.
In contrast to MSAs, certain drugs reduce MT polymerization, such as colchicine, vincristine (VNC), and nocodazole (NOC), and interact with free tubulin subunits, decreasing the concentration of free tubulin available to participate in MT dynamics, thus shifting the balance between polymerized and free subunits toward depolymerization and MT net mass loss [149,150,151,152] and incrementing the proportion of labile MTs, which also impair brain function [55]. NOC can increase MT dynamics in rats after spared nerve injury, improving their cognitive function [153]. Subsequently, describing the characteristics of MT and their endogenous and exogenous modulators, we will describe physiological and pathological conditions that modulate MT configuration. After, we will describe the opposite phenomenon: how changes in MT configuration impact neural network function and morphology.

4. Microtubular Reconfiguration during Brain Function

The most convincing evidence of MT modulation associated with specific brain functions and states occurs during the reconfiguration processes underlying learning and memory [55,154,155,156]. For instance, there is an increase in the amount of MTs [154], and MT turnover [55], associated with training in different memory tasks. Memory consolidation is accompanied by increased expression of MAP2 and β-tubulin [155,156,157,158], which is reflected in increased MAP2 immunohistochemical staining [159,160,161]. However, the changes in MAPs expression during learning and memory do not always comprise more protein synthesis, as opposite changes have also been found [162]. For instance, both Tau long isoforms expression and Tau-dependent P13K signaling are decreased during cocaine-associated memory formation, which was abolished by Tau overexpression [162]; when extinction is achieved in this paradigm, Tau long isoforms return to basal levels [162]. It is possible that learning-dependent expression changes in these Tau isoforms drive the dynamic state of neuronal MTs [163]. Indeed, activity-dependent Tau post-translational modifications that could affect their interactions with MTs have also been described [92,164]. As will be reviewed in detail, learning and memory can induce a biphasic change in MT stability (measured by tyrosinated tubulin levels) in a stathmin-dependent manner [55]. Changes in MTs have also been associated with memory deficits [165,166]. For example, social isolation and injuries caused by induced cerebral hypoperfusion result in memory deficits that correlate with decreased levels of α-tubulin and MAP2 [165,166] The cognitive impairment in senescence-accelerated (SAMP10) mice also correlates with a reduction in MAP2 and a simplification of the dendritic arbor [167]. Interestingly, as already mentioned, memory deficits in spared nerve injury rats increased the levels of stable MT, reflected in α-tubulin hyperacetylation, which can be reversed by the MT destabilizer NOC [153]. The cellular mechanisms behind the brain function-induced MT reconfiguration are diverse and will be reviewed next.

5. Microtubular Reconfiguration during Neuronal Activity

The generalized increase in neuronal activity, induced by KCl depolarization, is enough to induce MT polymerization, which is blocked by inhibiting action potentials generation with the Na+ channel blocker tetrodotoxin [168]. KCl-induced depolarization increases α-tubulin acetylation [169] and MT entry into dendritic spines [168]. Recently, chemogenetic activation of adult dorsal root ganglion neurons increased MT dynamics through tubulin acetylation, which resulted in axonal growth after nerve injury in vitro [170]. KCl-induced depolarization not only increases neuronal firing [171] but also releases neurotransmitters [172,173], including glutamate, GABA, and glycine [172,173], which modulate MT density [52,174,175,176,177,178]. For instance, activation of glutamate receptors modulates MT function by regulating the expression of MAP2 [52,175,176,177,178] and the upregulation of Tau translation and its accumulation in the somatodendritic compartments [179]. As will be reviewed later, NMDA receptor-dependent synaptic activation increases the proportion of dendritic spines containing dynamic MTs, contributing to spine morphological changes [53,168,180,181,182]. NMDA-dependent MT modulation also occurs during the induction and maintenance of both LTP [168,183] and long-term depression (LTD) [184], which have been related to changes in MT dynamicity in an EB3- and MAP2-dependent manner [184]. Most of the changes in MTs described so far could be explained, or at least partially influenced, by Ca2+ entry through NMDA receptors [182]. In vitro studies have shown that MTs polymerize at low Ca2+ concentrations, whereas MTs disassemble at increased Ca2+ concentrations [185,186,187]. These effects have been hypothesized to be mediated through direct interactions of Ca2+ with tubulins or, indirectly, by Ca2+-dependent regulators of MT assembly, such as calmodulin [188,189,190,191], by Ca2+-dependent modulation of MAPs (i.e., tau) [192] or by drebrin [182]. Similar changes in MTs can be induced by blocking glycinergic receptors, which facilitates tubulin polyglutamylation and alters binding of MAP2 to MTs, which is accompanied by reduced motor protein mobility and cargo delivery into neurites [193]. Reduction in synaptic inhibition has been associated with the induction of hyperexcitable states related to epilepsy [4,174]. Interestingly, the levels of Tyr-tubulin and MT dynamicity are dysregulated in both patients with intractable temporal lobe epilepsy and chronic models of epilepsy [194], while drugs that affect MT stability can either increase (i.e., colchicine) or decrease (i.e., noscapine) epileptiform activity [194]. In vitro experiments have shown that optogenetic neuronal stimulation promotes Tau release, which is reproduced in vivo along with the neuron-to-neuron spreading of this MAP in its pathological forms [195,196,197,198]. Next, we will review the influence of endogenous and exogenous MT modulators on neuronal excitability and morphology [196].

6. Changes in Excitability and Synaptic Transmission, and Their Morphological Correlates, Induced by Microtubular Reconfiguration

MT polymerization and depolymerization participate in the clustering and stabilization of ion channels [199], including Na+ channels [200], transient receptor potential (TRP) channels [201] and Ca2+ channels, and influence their functionality [202,203,204]. Thus, it is expected that endogenous MT stability modulators have diverse effects on neuronal excitability [75]. For instance, the absence of Tau protein makes animals resistant to seizures [205,206], while the reduction of Tau expression alleviates seizure burden and improves survival in some genetic models of epilepsy [206,207,208]. Altogether, these findings indicate that either absence or hyperphosphorylation of Tau produce a hypoexcitable state, which correlates with a reduction in neuronal firing and changes in neurotransmitter release probability [209]. However, this seems to not always be the case, since very young AD transgenic mice, expressing a mutant version of tau and already accumulating hyperphosphorylated Tau protein, do not produce seizure-like activity in the presence of the potassium channel blocker 4-aminopyridine [4]. Tau protein can induce changes in Kv4.2 expression dendrites of CA1 pyramidal cells, which alters their excitability and synaptic plasticity [210], as Tau transgenic animals become older, their neurons exhibit depolarized neuronal resting membrane potentials [207,208,209], increased evoked action potential firing [211,212,213] and are more prone to induced epilepsy [214,215]; a finding consistent with studies demonstrating increased seizure prevalence in patients with AD [191,216,217,218,219,220,221,222]. However, there is a report showing that the Tg4510 Tau mouse model, at similar ages, exhibits neurons with reduced action potential frequency [220,221]. Indeed, the rTg4510 transgenic line, which expresses 14-fold mutant Tau, compared to endogenous Tau, exhibits overall cortical hypoactivity [222]. These diverse changes in firing could be explained by tau-induced modulation of Na+ channel function [223,224] in different neuronal types and compartments [224,225,226,227,228]. Another endogenous MT modulator that influences Na+ channel function is FGF13 [229]. In fact, Fgf13 knockout mice show markers of MT instability correlating with a reduction of Na+ channel presence at the cell membrane, which is mimicked by colchicine in wild-type mice [229]. In contrast, FGF13 overexpression or PTX application results in more Na+ channel proteins being inserted into the surface membrane [229]. The latter finding indicates that exogenous MT modulators can affect neuronal activity, as will be reviewed next.
We have just mentioned that PTX promotes Na+ channel insertion into the surface membrane [229], which correlates with PTX-induced increased levels of endogenous Nav1.7 mRNA levels and Na+ current density [230]. Thus, the changes in ion channels induced by PTX [229,230] could be the cause of the non-convulsive status epilepticus, revealed by EEG, induced by chemotherapeutic administration of PTX [231]. This finding is similar to the slight increase in the frequency and duration of epileptiform activity induced by PTX in vitro [232]. In contrast, MT destabilization with NOC decreased or even abolished epileptiform activity in vitro, which was corroborated in vivo using the maximal dentate activation model [232]. This finding contrasts with the increase of spontaneous seizures during chronic epilepsy induced by colchicine, which is based on the decrease in interneuron firing and the reduction of their inhibitory postsynaptic currents [194]. In contrast, the MT-modulating agent noscapine increased the frequency of action potentials in interneurons and boosted their inhibitory postsynaptic currents, halting the progression of spontaneous seizure during chronic epilepsy [194]. In addition, the MT destabilizer VNC produces a complex effect on excitability since it enhances the excitability of some neurons while reducing it in others [225]. Thus, it appears that MT-modulating agents have heterogeneous and complex net effects depending on the molecule family and neuronal type they are tested on.
One cellular compartment that is highly modulated by MTs and has an enormous impact on neuronal excitability is the proximal region of the axon called axon initial segment (AIS) [221,233], which is a specialized compartment that has a high density of voltage-gated ion channels and generates action potentials [234,235,236,237]. This specialized region consists of the proximal portion along the first 20–40 μm of the axon that extends from the axon hillock to the beginning of the myelin sheath [238]. The AIS has a unique cytoskeletal organization compromising cytoskeletal submembrane networks [239,240,241]. These networks consist of MT bundles coated with a dense submembrane protein network containing ankyrin G (AnkG), βIV-spectrin, and actin filaments [239,242], which serve as scaffolds for ion channel localization and maintenance on the membrane. The AIS cytoskeleton forms a transport barrier between the axon and the somatodendritic membrane [243] and regulates axonal entry of cargoes that require selective transport [244]. The AIS also plays a key role in maintaining the molecular and functional neuronal polarity by controlling membrane diffusion and the polarized trafficking of cytoplasmic proteins toward the axon [244,245]. For example, MAP2 is specifically located in the somatodendritic region and this exclusion from the axonal compartment depends on the assembly of the AIS [241]. The AIS is particularly enriched in voltage-gated sodium (Na+) and potassium (K+) channels that are required for action potential generation, and the membrane-adaptor protein AnkG is the main component of the AIS scaffold that determines the functional and structural properties of the proteins located at the AIS by recruiting and clustering them in this region [246,247].
The AIS is an essential compartment in the integration of the excitatory and inhibitory postsynaptic potentials into action potential generation [248,249]. Components of the AIS, including the cytoskeleton and ion channels, undergo activity-dependent structural changes that modulate neuronal excitability and maintain steady-state firing rates [234,235,236,237,249]. For example, the AIS is elongated in avian neurons deprived of synaptic inputs for several days [236]. In response to chronic depolarization, the AIS location shifts distally in murine excitatory neurons [233,234,250], but proximally in inhibitory interneurons [251]. Functional defects of the AIS due to cytoskeletal alterations have been reported in cellular models of tauopathies [224,227,228,252]. Moreover, pathogenic Tau acetylation destabilizes the cytoskeletal submembrane protein AnkG and the MTs in the AIS [253], changes the location of the AIS [221,253], precludes activity-dependent change in AIS location [233], and reduces excitability [221]. Similarly, MT destabilization with NOC reproduces most of these effects on AIS and excitability [233]. In contrast, stabilizing MTs with Epo-D restores the cytoskeletal barrier in the AIS and prevents Tau mislocalization [253]. Moreover, PTX prevents pathogenic Tau-induced AIS mislocalization and normalizes neuronal excitability [233]. Considering that neuronal excitability is only a part of the basic processes underlying neural network activity [254], next we will review the changes in synaptic activity induced by endogenous and exogenous MT modulators.
At the presynaptic compartments, the cytoskeleton enters deep into presynaptic terminal swellings and partially colocalizes with a subset of synaptic vesicles (SVs) [255]. Functionally, it is thought that this interaction regulates neurotransmitter release [255,256], mediates endocytosis of SVs [257,258,259], and promotes the recovery of synaptic responses from activity-dependent short-term depression [260,261] via fast SV replenishment [262], clearance of used SVs from release sites [263,264,265,266], and transport of mitochondria and presynaptic elements [266,267]. Early electron microscopy studies at the frog neuromuscular junction reported that MTs anchoring SVs are directed toward active zones [268,269,270]. Likewise, at the Drosophila NMJ, the MT-associated protein Futsch [271] links MTs to AZs, thereby supporting neurotransmitter release [256]. At the calyx of Held in adult cats, MTs are observed in presynaptic terminal swellings, but not in the SV pool [272]. Depolymerization of MTs with NOC impairs long-distance SV movements between presynaptic swellings [273]. Dynamic MTs preferentially growing in presynaptic boutons show biased directionality in that they are almost always oriented toward the distal tip of the axon, which can be modulated by neuronal activity [23]. Silencing γ-tubulin expression reduces presynaptic MT nucleation, SV interbouton transport and regulates evoked SV exocytosis [23]. Dynamic MTs are enriched at en passant boutons and allow for the targeted delivery and unloading of SV precursors by the kinesin-3 motor KIF1A [274]. In en passant boutons, presynaptic dynamic MTs are nucleated upon neuronal activity and are critical for adjusting activity-evoked neurotransmitter release by providing paths for interbouton bidirectional transport of SVs, which is a rate-limiting step in SV unloading and exocytosis at release sites [23,274]. MT nucleation preferentially occurs at excitatory boutons in hippocampal slices from neonatal mice [23]. Dynamic MTs may be directly regulating Ca2+ handling at terminals through the interaction of EB1/3 with endoplasmic reticulum Ca2+ sensors [275,276,277]. In simultaneous presynaptic and postsynaptic action potential recordings, depolymerization of MTs impaired the fidelity of high-frequency neurotransmission at the calyx of Held presynaptic terminals [24]. Thus, it would be expected that modifications in MTs would have a major impact in synaptic transmission.
As shown for neuronal excitability, changes in Tau also affect basal synaptic activity [209,211,212,213,278,279]. As mentioned, Tau knockout changes synaptic release probability which is reflected in an increase in paired-pulse facilitation [209]. In contrast, the overexpression of mutated human Tau increases spontaneous excitatory postsynaptic currents (EPSPs) [211,212,213], increases glutamate release, and decreases glutamate reuptake [278,279], while also decreasing paired-pulse facilitation [280,281,282,283]. MT influence on synaptic transmission has been observed after the depletion of the MT-severing protein spastin, which produces longer MTs with increased tubulin polyglutamylation leading to a lower frequency of miniature EPSCs [284]. Beyond these effects of endogenous MT modulators on synaptic transmission, there is extensive evidence of the effects of pharmacological MT modulators on this phenomenon. For instance, MT stabilization with PTX increases the frequency of miniature EPSCs and reduces the paired-pulse facilitation of evoked EPSCs, which is reversed by the NMDA receptor antagonist 2-amino-5-phosphonopentanoic acid [285]. This is similar to the increase in amplitude and frequency of miniature inhibitory postsynaptic currents induced by noscapine [194]. These observations contrast with the effects observed by MT destabilization with colchicine, which reduces the amplitude and mIPSCs [194], similarly to the reduction in synaptic transmission recovery in the presence of vinblastine at the calyx of Held presynaptic terminals [24]. As mentioned, in these highly active synapses the presynaptic MTs play important roles in SV cycling and mitochondrial anchoring [24,273,286]. Although we will later discuss the influence of MTs on dendritic spines, it is important to conclude this section by indicating that NMDA and AMPA receptors trafficking, depends heavily on stable MT-mediated transport [57,287,288]. For instance, NOC or colchicine agents inhibited NMDA receptor-mediated function and, thus, synaptic currents in an MT-dependent manner [287]. The effect of MT depolymerizers, which is most prominent in NR2B subunit-containing NMDA receptors, was blocked by cellular knockdown of the kinesin motor protein KIF17, which transports NR2B-containing vesicles along MTs in neuronal dendrites. Moreover, immunocytochemical studies show that MT depolymerizers decreased the number of surfaces NR2B subunits on dendrites, all of which were reversed by brain-derived neurotrophic factor (BDNF) and PTX [289]. To further support the role of MTs in synaptic receptors, memory consolidation regulates, in a stathmin-dependent manner, the transport of the GluA2 subunit of the AMPA receptor, resulting in increased GluA2 at synaptic sites, which promotes long-term memory [56]. The complex relationship between MTs and memory will be described next.

7. Changes in Long-Term Synaptic Plasticity and in Dendritic Spines Induced by Microtubular Modulation

Endogenous and pharmacological MT modulation has a major impact on learning and memory [55,56]. To understand this impact, we will first review the effects of endogenous and pharmacological MT modulation on two of its most likely underlying cellular mechanisms, namely long-term synaptic plasticity [290] and dendritic spine reconfiguration [1,53,168,182,187,291,292]. There is extensive evidence that changes in MTs strongly affect long-term plasticity [55,56,293,294]. For instance, LTP is highly dependent on normal Tau function, in a very narrow homeostatic range, since this process is abolished by either Tau knockout [295] or the overexpression of pathological forms of Tau [278,280,281,282,283,296]. Pathological Tau constructs reduce LTP in CA3-CA1 connection [296,297], while LTD is altered by the presence of hyperphosphorylated Tau due to changes in NMDA receptor activity [164,298]. In contrast, it is also reported that young transgenic mice expressing a mutant form of Tau exhibit an increase in LTP [299], while LTD can be inhibited by very low concentrations of oligomeric Tau [300]. Another MT modulator that plays a major role in LTP regulation is stathmin, a protein that binds tubulin and inhibits MT polymerization [56,293]. Mice lacking the stathmin4A isoform or its non-phosphorylatable mutant exhibit deficits in LTP generation in the cortico-amygdala, thalamo-amygdala, and the perforant path to the dentate gyrus synapses, but not at the Schaffer collaterals to CA1 [56,293]. Spastin depletion also reduces LTP [284]. Altogether, these findings indicate that normal MT function is required for the induction of synaptic potentiation, while the use of exogenous MT modulators reveals a similar scenario. Slices treated with NOC cannot maintain post-tetanic potentiation or LTP [53,55,301], although other authors have not found this decline in the presence of VNC [302]. However, synaptic potentiation in the presence of VNC becomes sensitive to the co-application of protein synthesis inhibitors [301]. Similarly, MT stabilization with PTX also reduces LTP in the cortico-amygdala and CA3-CA1 synapses [149,293]. However, Epo-D can reestablish LTP in animals lacking MAP6, which were unable to induce such potentiation, at the CA3-CA1 synapse [135]. As for LTP, LTD is highly sensitive to MT modulation [303,304]. Hippocampal LTD deficit is common in Tau knockouts [303,304] or mice with a reduced expression of Tau [304]. Nonetheless, Epo-D does not affect LTD induction at the Schaffer collaterals reaching the CA1 [135].
Dendritic spines are small micrometer-sized specialized protrusions of the membrane that decorate dendritic branches [305] and act as dynamic microcompartments to restrict and amplify excitatory signaling [306] and whose plasticity has been associated with a variety of neural functions, including learning and memory [166,181]. Spine shape and function were classically considered to be mainly determined by actin filaments, which are highly enriched in spines [307]. However, now it is well accepted that MTs are determinant in the development, maintenance, plasticity, and degeneration of spines [53,308,309,310,311,312]. For instance, pR5 transgenic mice, which overexpress hyperphosphorylated Tau, show significant changes in dendritic morphology in CA1 pyramidal neurons [313]. Moreover, dynamic MTs appear to regulate dendritic spine morphology and synaptic plasticity [53] and promote NMDA receptor and Ca2+-dependent spine enlargement [181] by continuously invading all types of spines (mushroom, stubby and thin, as well as filopodia). Large spines consistently exhibit transient and activity-dependent invasion of MTs [168,314], which mainly relies on Ca2+ [168,180,182], membrane depolarization [168] and the interaction of MTs with actin through drebrin [178,315,316]. MT invasion of spines is also highly dependent on the end-binding protein 3 (EB3) [53]. In fact, inhibition of MT growth by depletion of EB3 caused the specific loss of mushroom-headed spines and increased the percentage of filopodia [53,314]. As expected, EB3 overexpression causes the increase of mushroom-headed spines [53,316]. Moreover, EB3 overexpression reverses the deficiency of mushroom spines in AD transgenic mice [317].
Spine invasion by MTs is a transient event [53,168,181,182,184,291,292] that occurs under physiological conditions [53,168,312] but that is exacerbated with stimulations that induce LTP [168,182,318,319], with membrane depolarization [168] or with the application of BDNF [320]. NMDA receptor-dependent synaptic activation increased the proportion of dendritic spines containing dynamic MTs, which then contributed to spine enlargement [53,168,180,182]. On the other hand, inhibition of NMDA receptor activity reduced MT invasion of spines [181]. In contrast, stimulation of both synaptic and extrasynaptic NMDA receptors by bath application of NMDA results in a loss of MT dynamics in dendrites and spines [184], inducing LTD [184], which requires removing EB3 from the growing MTs in a Ca2+-dependent manner [184]. LTP and high KCl-induced increase of dendritic spines containing MTs were completely abolished by inhibiting the firing of action potentials with TTX [168]. Long-term treatment of hippocampal cultures with BDNF increases spine number, which is further increased by the presence of the MT-stabilizing agent PTX [314]. In contrast, disruption of MTs with NOC blocks the spine-promoting effect of BDNF [314]. Importantly, MT entry into spines was increased after the transient stimulation with KCl, and this increase was blocked by treatment with TTX, indicating that MT dynamics in neurons are changing in an action potential-dependent manner [168].
Neuronal MAPs, such as MAP2 and Tau are also involved in regulating MT dynamics and interactions in dendritic spines [321,322]. MAP2 binds along the length of MTs but is also associated with actin in dendritic spines [319] and interacts with the NMDA receptor subunits NR2A and NR2B [322]. Similar to MAP2, Tau interacts with PSD-95, which in turn regulates NMDA receptor through the tyrosine kinase Fyn [164]. Therefore, NMDA receptor-mediated Tau phosphorylation at specific residues results in the weakening of the tau–PSD95–Fyn interaction, regulating postsynaptic plasticity [5,164]. Dynamic MT spine invasion regulates spine morphology [93,323,324], synaptic plasticity [168,183], the recycling of endosomes containing AMPA receptors into spines from the dendritic shaft [325] and the content of PSD95 in the spines [320]. Syntaptotagmin4-containing vesicles are also transported by polymerizing MTs into spine heads, where they subsequently undergo exocytosis [292]. Thus, it is not surprising that shortening of dendritic spines and changes in spine shape (i.e., shift from mushroom to stubby spines) appear to be relevant indicators of the progression of cognitive deficits [92,323,324] and that these and other alterations of spines have been reported in different brain pathologies such as autism spectrum disorders, schizophrenia, and fragile X syndrome [69,326,327].
Exogenous MT modulators have a huge impact on spine shape and function. For instance, NOC causes spine loss without changing spine morphology [48]. In another study, NOC abolished EB3 accumulation at MTs and reduced the number of mushroom spines while increasing the number of filopodia [53]. In addition, NOC suppresses the spine recovery induced by a gamma-secretase inhibitor in AD transgenic mice [48]. NOC also reversed the memory-induced increase in MAP2-associated MTs, reducing dendritic spine density, and impairing memory formation. The effects of NOC on MT turnover were prevented by PTX and BDNF, which restored dendritic spine density and memory formation [55]. Similarly, NOC inhibits the BDNF-induced increase in the spine density, yet [314] it also partially restores the number of mushroom spines and spine density in AD transgenic neurons, possibly by promoting dynamic MT entry into the spines [48]. Similarly, Epo-D also reverses Aβ-induced spine loss [328], which would appear to be counterintuitive, as Epo-D reduces mushroom spines in wild-type slice cultures [328]. Thus, it is possible that Epo-D prevents MT disassembly in AD transgenic neurons and maintains spine morphology; however, it also inhibits MT dynamics required for spine maturation in normal mice [328]. In fact, other groups have found that while Epo-D alters dendritic spine length, density, and morphology [136,329,330], it can also reduce spine length and increase the density of mushroom spines after fluid percussion injury [140]. PTX also reduces dendritic spine density, which is mitigated in Tau knockout neurons [331]. Although PTX alters the dynamics of dendritic spines [332], it also prevents the reduction of spine density and memory alterations induced by NOC [55] and exacerbates the BDNF-induced increase in the spine density [314].

8. Changes in Learning and Memory Induced by Microtubular Modulation

Learning and memory require the proper representations of experiences that become imprinted in neuronal circuits during memory consolidation [290], which would involve functional and morphological changes that depend on MT function. Thus, it is not surprising that one of the most challenging side-effects of MSA-based chemotherapy is learning and memory impairments [130,333,334,335,336]. However, as will be reviewed next, MT instability or stability can either promote or impede learning and memory in a state-dependent manner [56,57]. The clearest example of this dynamic relationship was provided by Uchida et al. (2014) and Yousefzadeh et al., (2021) [56,337] who found that learning and memory cause biphasic changes in MTs. In the early phase of the process, stathmin dephosphorylation enhances MT-instability, whereas in the late phase these processes are reversed and a hyperstable MT state is achieved during context-fear memory [56]. As expected, PTX administration immediately following the training precludes memory formation [56,337] but increases memory when applied during its maintenance [56,337].
The complex interaction between MTs and learning and memory is well exemplified by the diverse actions that Tau exerts on MT. For instance, the expression of Tau mutant variants in Drosophila not only alters the cytoskeleton at the synaptic terminals but also modifies neuronal activity patterns and memory consolidation [338]. Similarly, normal Tau overexpression results in learning and memory deficits [339,340]. Moreover, old transgenic mice expressing mutant forms of Tau also exhibit deteriorated memory [54,116,143] and disrupted LTP [278,280,281,282,283,296]. However, young transgenic mice expressing a mutant form of Tau exhibit improved memory [299], which correlates with increased LTP [299]. Changes in other MT regulators also affect learning and memory in a complex manner. In the case of stathmin, its knockout impaired memory and LTP [293], while the expression of the non-phosphorylatable and constitutively active form produces similar effects [58]. Other examples of learning and memory alterations induced by the reduction of MT stabilizers have been found in mice lacking FGF13 [75], spastin [286], CRTC1 [341], or KIF21B [342].
The effects of exogenous modulators of MT stability on learning and memory are also complex. Epo-D may have some neuroprotective effects on this phenomenon since its application improves cognitive performance in Tau transgenic mice [54,112,139,343] by increasing MT density and axonal integrity and decreasing hyperdynamic MTs [54,112,139,343]. Similarly, EpoD treatment has beneficial effects on APP/PS1 double-transgenic mice, improving their axonal transport of mitochondria-associated with enhanced motor and spatial memory [344]. However, Epo-D induces an alteration in reversal learning (crossover) in these animals [344]. Similarly, PTX can prevent traumatic brain injury-induced deficits in memory [345], which is due to the prevention of structural injury and hypometabolism [345]. PTX also prevents the memory impairment induced by NOC [55]. However, as already mentioned, PTX can have a deleterious effect on learning and memory [56], which has been observed by many groups [143,153,346,347,348,349,350,351] suggesting that state-dependent physiological MT dynamics, rather than an overall shift to stabilization, is important for learning and memory. Thus, since a moderate stabilization of MTs may be protective, prevention of MT dynamics can have a detrimental effect on these plastic phenomena [154,346,347,348,349,350,351,352,353]. In fact, under conditions in which memory was not impaired, PTX treatment impaired learning of new rules [59]. PTX-induced memory impairment, which can be prevented by lithium [351], has been related to a decrease in LTP and [154], neurogenesis [351], an increased number of TUNEL-positive neurons, increased expression of TNF-α and IL-1β [352] and can be reduced by the TNF-α synthesis inhibitor thalidomide [353], which indicates that this phenomenon could be related to neuroinflammation [353]. However, it is important to notice that others have found that PTX does not induce brain inflammation, as measured by cytokine analysis, which correlates with the lack of effect of aspirin on PTX-induced memory alteration [346]. PTX-induced memory impairment has also been associated with a reduction in dendritic length and complexity [354], which is also reverted by lithium [354]. This impairment has also been related to a reduction in cell proliferation [350,351]. As described for PTX, DTX also induces alterations in memory [347,348,355] which, in this case, is related to elevated neural autophagy and astrocytic activation [352] and is reversed by rolipram administration [50,348].
The clearest evidence that MT destabilization affects memory is provided by extensive demonstrations that colchicine affects learning and memory in an MT-dependent manner [355,356,357,358,359]. However, it should be acknowledged that, at high doses, colchicine can induce cell death [360,361,362,363,364,365,366]. Although MT destabilization with NOC administration immediately following training could promote learning [55,336], NOC also can inhibit memory formation if administered before learning [54] and even reduce memory retrieval if administered at late phases of learning [55,336]. However, NOC can also prevent the memory deficits induced by spared nerve injury [152].
In addition to the modulation of encoding and retrieval, some effects of MT stabilizers could also be interpreted as impairments on cognitive flexibility, which comprehends adaptative changes in the behavioral output in response to modifications of the rules of the task [367,368]. For instance, weekly administration of Epo-D not only prevented APP/PS1 mice from exhibiting retrieval impairments in the water Morris maze [344], but also caused less extinction, that is expected when the platform is removed for several weeks [344]. Moreover, the injection of PTX in the dorsolateral striatum, after the learning of a specific target duration in a temporal learning paradigm, prevents the acquisition of a new target duration but strengthens the recall of the old one [337]. This rigid behavior also accounts for reduced cognitive flexibility [367,368]. An elegant work showed that PTX has negative effects specifically on reversal learning, sparing prior training with simple discrimination of pairs of odorants and the learning of new pairs of odorants [350]. Moreover, a single dose of PTX to cancer patients has been associated with confusion, and word recollection impairments [130,336]. On the other hand, MT destabilization can also impair cognitive flexibility. Stat4A mice have reduced reversal learning in the Morris water maze [57], which comprehends the relocation of the hidden platform after the training with the original location [57]. These effects of MT modulation on cognitive flexibility could be due to the impairments induced by MT stabilizers and destabilizers on brain connectivity [126]. MT stabilizers-induced neuropathic pain [122,123] could also reduce cognitive flexibility as observed in some pain models [369].
Overall, it appears that either endogenous homeostatic and pathological MTs modulation, or exogenous pharmacological modulation of MT dynamicity, encompasses favorable and detrimental effects in brain shape and function. Some factors like age, sex, brain region, neuronal type, learning, activity-dependent processes, and behavioral tasks influence the dynamic state of neuronal MTs and bias the effects of some stabilizing and destabilizing drugs. This could be due to the exquisitely regulated allostasis that MTs exhibit to respond effectively to specific neuronal demands in their cytosolic microenvironment. Further research must be focused on the underlying mechanisms of MT-dependent processes that modify behavior with the more temporal and temporal resolution, considering age, sex, and strain as key factors. Thus, some basic gaps to fill are the sex, brain region, and neuronal type dependence of the MT modulation and its effects on brain shape and function. It would be relevant to also investigate the protein-protein interactions and post-translational modifications occurring during MT dynamic response to neuronal function.

Author Contributions

F.P.-O., Á.A.R.-G. and L.X.-C. reviewed the literature and integrated the information required for this review. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by Dirección General de Asuntos del Personal Académico UNAM (Grant IG200521) and CONACyT (Grant A1-S-7540).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We would like to thank Benito Ordaz for technical help, Jessica González and Anaid Antaramian for editing the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kapitein, L.C.; Hoogenraad, C.C. Which Way to Go? Cytoskeletal Organization and Polarized Transport in Neurons. Mol. Cell. Neurosci. 2011, 46, 9–20. [Google Scholar] [CrossRef] [PubMed]
  2. Witte, H.; Bradke, F. The Role of the Cytoskeleton during Neuronal Polarization. Curr. Opin. Neurobiol. 2008, 18, 479–487. [Google Scholar] [CrossRef] [PubMed]
  3. Solís-Chagoyán, H.; Calixto, E.; Figueroa, A.; Montaño, L.M.; Berlanga, C.; Rodríguez-Verdugo, M.S.; Romo, F.; Jiménez, M.; Gurrola, C.Z.; Riquelme, A.; et al. Microtubule Organization and L-Type Voltage-Activated Calcium Current in Olfactory Neuronal Cells Obtained from Patients with Schizophrenia and Bipolar Disorder. Schizophr. Res. 2013, 143, 384–389. [Google Scholar] [CrossRef] [PubMed]
  4. Mondragón-Rodríguez, S.; Gu, N.; Manseau, F.; Williams, S. Alzheimer’s Transgenic Model Is Characterized by Very Early Brain Network Alterations and β-CTF Fragment Accumulation: Reversal by β-Secretase Inhibition. Front. Cell. Neurosci. 2018, 12, 121. [Google Scholar] [CrossRef]
  5. Mondragón-Rodríguez, S.; Salgado-Burgos, H.; Peña-Ortega, F. Circuitry and Synaptic Dysfunction in Alzheimer’s Disease: A New Tau Hypothesis. Neural Plast. 2020, 2020, 2960343. [Google Scholar] [CrossRef]
  6. Dent, E.W.; Merriam, E.B.; Hu, X. The Dynamic Cytoskeleton: Backbone of Dendritic Spine Plasticity. Curr. Opin. Neurobiol. 2011, 21, 175–181. [Google Scholar] [CrossRef] [Green Version]
  7. Baas, P.W.; Rao, A.N.; Matamoros, A.J.; Leo, L. Stability Properties of Neuronal Microtubules. Cytoskeleton 2016, 73, 442–460. [Google Scholar] [CrossRef] [Green Version]
  8. Sánchez-Huertas, C.; Freixo, F.; Viais, R.; Lacasa, C.; Soriano, E.; Lüders, J. Non-Centrosomal Nucleation Mediated by Augmin Organizes Microtubules in Post-Mitotic Neurons and Controls Axonal Microtubule Polarity. Nat. Commun. 2016, 7, 12187. [Google Scholar] [CrossRef]
  9. Nogales, E.; Wang, H.-W. Structural Intermediates in Microtubule Assembly and Disassembly: How and Why? Curr. Opin. Cell Biol. 2006, 18, 179–184. [Google Scholar] [CrossRef]
  10. Weisenberg, R.; Taylor, E.W. Studies on ATPase Activity of Sea Urchin Eggs and the Isolated Mitotic Apparatus. Exp. Cell Res. 1968, 53, 372–384. [Google Scholar] [CrossRef]
  11. Bryan, J.; Wilson, L. Are Cytoplasmic Microtubules Heteropolymers? Proc. Natl. Acad. Sci. USA 1971, 68, 1762–1766. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Walker, R.A.; O’Brien, E.T.; Pryer, N.K.; Soboeiro, M.F.; Voter, W.A.; Erickson, H.P.; Salmon, E.D. Dynamic Instability of Individual Microtubules Analyzed by Video Light Microscopy: Rate Constants and Transition Frequencies. J. Cell Biol. 1988, 107, 1437–1448. [Google Scholar] [CrossRef] [PubMed]
  13. Burns, R.G.; Farrell, K.W. Getting to the Heart of β-Tubulin. Trends Cell Biol. 1996, 6, 297–303. [Google Scholar] [CrossRef]
  14. Mejillano, M.R.; Barton, J.S.; Nath, J.P.; Himes, R.H. GTP Analogs Interact with the Tubulin Exchangeable Site during Assembly and upon Binding. Biochemistry 1990, 29, 1208–1216. [Google Scholar] [CrossRef]
  15. Weisenberg, R.C.; Deery, W.J.; Dickinson, P.J. Tubulin-Nucleotide Interactions during the Polymerization and Depolymerization of Microtubules. Biochemistry 1976, 15, 4248–4254. [Google Scholar] [CrossRef]
  16. David-Pfeuty, T.; Erickson, H.P.; Pantaloni, D. Guanosinetriphosphatase Activity of Tubulin Associated with Microtubule Assembly. Proc. Natl. Acad. Sci. USA 1977, 74, 5372–5376. [Google Scholar] [CrossRef] [Green Version]
  17. Yu, W.; Centonze, V.; Ahmad, F.; Baas, P. Microtubule Nucleation and Release from the Neuronal Centrosome. J. Cell Biol. 1993, 122, 349–359. [Google Scholar] [CrossRef]
  18. Kapitein, L.C.; Hoogenraad, C.C. Building the Neuronal Microtubule Cytoskeleton. Neuron 2015, 87, 492–506. [Google Scholar] [CrossRef] [Green Version]
  19. Ahmad, F.J.; Baas, P.W. Microtubules Released from the Neuronal Centrosome Are Transported into the Axon. J. Cell Sci. 1995, 108, 2761–2769. [Google Scholar] [CrossRef]
  20. Dent, E.W. Of Microtubules and Memory: Implications for Microtubule Dynamics in Dendrites and Spines. Mol. Biol. Cell 2017, 28, 1–8. [Google Scholar] [CrossRef]
  21. Heidemann, S.R.; Landers, J.M.; Hamborg, M.A. Polarity Orientation of Axonal Microtubules. J. Cell Biol. 1981, 91, 661–665. [Google Scholar] [CrossRef] [PubMed]
  22. Vale, R.D. The Molecular Motor Toolbox for Intracellular Transport. Cell 2003, 112, 467–480. [Google Scholar] [CrossRef] [Green Version]
  23. Qu, X.; Kumar, A.; Blockus, H.; Waites, C.; Bartolini, F. Activity-Dependent Nucleation of Dynamic Microtubules at Presynaptic Boutons Controls Neurotransmission. Curr. Biol. 2019, 29, 4231–4240.e5. [Google Scholar] [CrossRef] [PubMed]
  24. Piriya Ananda Babu, L.; Wang, H.-Y.; Eguchi, K.; Guillaud, L.; Takahashi, T. Microtubule and Actin Differentially Regulate Synaptic Vesicle Cycling to Maintain High-Frequency Neurotransmission. J. Neurosci. 2020, 40, 131–142. [Google Scholar] [CrossRef]
  25. Baas, P.W.; Deitch, J.S.; Black, M.M.; Banker, G.A. Polarity Orientation of Microtubules in Hippocampal Neurons: Uniformity in the Axon and Nonuniformity in the Dendrite. Proc. Natl. Acad. Sci. USA 1988, 85, 8335–8339. [Google Scholar] [CrossRef] [Green Version]
  26. Baas, P.W.; Slaughter, T.; Brown, A.; Black, M.M. Microtubule Dynamics in Axons and Dendrites. J. Neurosci. Res. 1991, 30, 134–153. [Google Scholar] [CrossRef]
  27. Qiang, L.; Sun, X.; Austin, T.O.; Muralidharan, H.; Jean, D.C.; Liu, M.; Yu, W.; Baas, P.W. Tau Does Not Stabilize Axonal Microtubules but Rather Enables Them to Have Long Labile Domains. Curr. Biol. 2018, 28, 2181–2189.e4. [Google Scholar] [CrossRef] [Green Version]
  28. Poulain, F.E.; Sobel, A. The Microtubule Network and Neuronal Morphogenesis: Dynamic and Coordinated Orchestration through Multiple Players. Mol. Cell. Neurosci. 2010, 43, 15–32. [Google Scholar] [CrossRef]
  29. Kirschner, M. Beyond Self-Assembly: From Microtubules to Morphogenesis. Cell 1986, 45, 329–342. [Google Scholar] [CrossRef]
  30. Mitchison, T.; Kirschner, M. Dynamic Instability of Microtubule Growth. Nature 1984, 312, 237–242. [Google Scholar] [CrossRef]
  31. Holy, T.E.; Leibler, S. Dynamic Instability of Microtubules as an Efficient Way to Search in Space. Proc. Natl. Acad. Sci. USA 1994, 91, 5682–5685. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Roll-Mecak, A. The Tubulin Code in Microtubule Dynamics and Information Encoding. Dev. Cell 2020, 54, 7–20. [Google Scholar] [CrossRef] [PubMed]
  33. Waites, C.; Qu, X.; Bartolini, F. The Synaptic Life of Microtubules. Curr. Opin. Neurobiol. 2021, 69, 113–123. [Google Scholar] [CrossRef] [PubMed]
  34. Marchisella, F.; Coffey, E.T.; Hollos, P. Microtubule and Microtubule Associated Protein Anomalies in Psychiatric Disease. Cytoskeleton 2016, 73, 596–611. [Google Scholar] [CrossRef] [Green Version]
  35. Verhey, K.J.; Gaertig, J. The Tubulin Code. Cell Cycle 2007, 6, 2152–2160. [Google Scholar] [CrossRef]
  36. Yu, I.; Garnham, C.P.; Roll-Mecak, A. Writing and Reading the Tubulin Code. J. Biol. Chem. 2015, 290, 17163–17172. [Google Scholar] [CrossRef] [Green Version]
  37. Janke, C.; Bulinski, J.C. Post-Translational Regulation of the Microtubule Cytoskeleton: Mechanisms and Functions. Nat. Rev. Mol. Cell Biol. 2011, 12, 773–786. [Google Scholar] [CrossRef]
  38. Fukushima, N.; Furuta, D.; Hidaka, Y.; Moriyama, R.; Tsujiuchi, T. Post-Translational Modifications of Tubulin in the Nervous System. J. Neurochem. 2009, 109, 683–693. [Google Scholar] [CrossRef] [PubMed]
  39. Dunn, S.; Morrison, E.E.; Liverpool, T.B.; Molina-París, C.; Cross, R.A.; Alonso, M.C.; Peckham, M. Differential Trafficking of Kif5c on Tyrosinated and Detyrosinated Microtubules in Live Cells. J. Cell Sci. 2008, 121, 1085–1095. [Google Scholar] [CrossRef] [Green Version]
  40. Konishi, Y.; Setou, M. Tubulin Tyrosination Navigates the Kinesin-1 Motor Domain to Axons. Nat. Neurosci. 2009, 12, 559–567. [Google Scholar] [CrossRef]
  41. Hammond, J.W.; Huang, C.-F.; Kaech, S.; Jacobson, C.; Banker, G.; Verhey, K.J. Posttranslational Modifications of Tubulin and the Polarized Transport of Kinesin-1 in Neurons. Mol. Biol. Cell 2010, 21, 572–583. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Sudo, H.; Baas, P.W. Acetylation of Microtubules Influences Their Sensitivity to Severing by Katanin in Neurons and Fibroblasts. J. Neurosci. 2010, 30, 7215–7226. [Google Scholar] [CrossRef] [PubMed]
  43. Wade, R.H. On and Around Microtubules: An Overview. Mol. Biotechnol. 2009, 43, 177–191. [Google Scholar] [CrossRef] [PubMed]
  44. Slaughter, T.; Black, M.M. STOP (Stable-Tubule-Only-Polypeptide) Is Preferentially Associated with the Stable Domain of Axonal Microtubules. J. Neurocytol. 2003, 32, 399–413. [Google Scholar] [CrossRef] [PubMed]
  45. Sánchez, C.; Dı́az-Nido, J.; Avila, J. Phosphorylation of Microtubule-Associated Protein 2 and Its Relevance for the Regulation of the Neuronal Cytoskeleton Function. Prog. Neurobiol. 2000, 61, 133–168. [Google Scholar] [CrossRef]
  46. Dehmelt, L.; Halpain, S. The MAP2/Tau Family of Microtubule-Associated Proteins. Genome Biol. 2004, 6, 204. [Google Scholar] [CrossRef] [Green Version]
  47. Schwenk, B.M.; Lang, C.M.; Hogl, S.; Tahirovic, S.; Orozco, D.; Rentzsch, K.; Lichtenthaler, S.F.; Hoogenraad, C.C.; Capell, A.; Haass, C.; et al. The FTLD Risk Factor TMEM106B and MAP6 Control Dendritic Trafficking of Lysosomes. EMBO J. 2013, 33, 450–467. [Google Scholar] [CrossRef]
  48. Penazzi, L.; Bakota, L.; Brandt, R. Microtubule Dynamics in Neuronal Development, Plasticity, and Neurodegeneration. Int. Rev. Cell Mol. Biol. 2016, 321, 89–169. [Google Scholar]
  49. Binder, L.I.; Frankfurter, A.; Rebhun, L.I. The Distribution of Tau in the Mammalian Central Nervous System. J. Cell Biol. 1985, 101, 1371–1378. [Google Scholar] [CrossRef] [Green Version]
  50. Matus, A. Microtubule-Associated Proteins and the Determination of Neuronal Form. J. Physiol. 1990, 84, 134–137. [Google Scholar]
  51. Vale, R.D.; Schnapp, B.J.; Mitchison, T.; Steuer, E.; Reese, T.S.; Sheetz, M.P. Different axoplasmic proteins generate movement in opposite directions along microtubules in vitro. Cell 1985, 43, 623–632. [Google Scholar] [CrossRef]
  52. Aoki, C.; Siekevitz, P. Ontogenetic Changes in the Cyclic Adenosine 3′,5′-Monophosphate- Stimulatable Phosphorylation of Cat Visual Cortex Proteins, Particularly of Microtubule-Associated Protein 2 (MAP 2): Effects of Normal and Dark Rearing and of the Exposure to Light. J. Neurosci. 1985, 5, 2465–2483. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Jaworski, J.; Kapitein, L.C.; Gouveia, S.M.; Dortland, B.R.; Wulf, P.S.; Grigoriev, I.; Camera, P.; Spangler, S.A.; di Stefano, P.; Demmers, J.; et al. Dynamic Microtubules Regulate Dendritic Spine Morphology and Synaptic Plasticity. Neuron 2009, 61, 85–100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Barten, D.M.; Fanara, P.; Andorfer, C.; Hoque, N.; Wong, P.Y.A.; Husted, K.H.; Cadelina, G.W.; DeCarr, L.B.; Yang, L.; Liu, V.; et al. Hyperdynamic Microtubules, Cognitive Deficits, and Pathology Are Improved in Tau Transgenic Mice with Low Doses of the Microtubule-Stabilizing Agent BMS-241027. J. Neurosci. 2012, 32, 7137–7145. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Fanara, P.; Husted, K.H.; Selle, K.; Wong, P.Y.A.; Banerjee, J.; Brandt, R.; Hellerstein, M.K. Changes in Microtubule Turnover Accompany Synaptic Plasticity and Memory Formation in Response to Contextual Fear Conditioning in Mice. Neuroscience 2010, 168, 167–178. [Google Scholar] [CrossRef]
  56. Uchida, S.; Martel, G.; Pavlowsky, A.; Takizawa, S.; Hevi, C.; Watanabe, Y.; Kandel, E.R.; Alarcon, J.M.; Shumyatsky, G.P. Learning-Induced and Stathmin-Dependent Changes in Microtubule Stability Are Critical for Memory and Disrupted in Ageing. Nat. Commun. 2014, 5, 4389. [Google Scholar] [CrossRef] [Green Version]
  57. Uchida, S.; Shumyatsky, G.P. Deceivingly Dynamic: Learning-Dependent Changes in Stathmin and Microtubules. Neurobiol. Learn. Mem. 2015, 124, 52–61. [Google Scholar] [CrossRef] [Green Version]
  58. Martel, G.; Uchida, S.; Hevi, C.; Chevere-Torres, I.; Fuentes, I.; Park, Y.J.; Hafeez, H.; Yamagata, H.; Watanabe, Y.; Shumyatsky, G.P. Genetic Demonstration of a Role for Stathmin in Adult Hippocampal Neurogenesis, Spinogenesis, and NMDA Receptor-Dependent Memory. J. Neurosci. 2016, 36, 1185–1202. [Google Scholar] [CrossRef] [Green Version]
  59. Smith, A.E.; Slivicki, R.A.; Hohmann, A.G.; Crystal, J.D. The Chemotherapeutic Agent Paclitaxel Selectively Impairs Learning While Sparing Source Memory and Spatial Memory. Behav. Brain Res. 2017, 320, 48–57. [Google Scholar] [CrossRef] [Green Version]
  60. Matamoros, A.J.; Baas, P.W. Microtubules in Health and Degenerative Disease of the Nervous System. Brain Res. Bull. 2016, 126, 217–225. [Google Scholar] [CrossRef] [Green Version]
  61. Brandt, R.; Bakota, L. Microtubule Dynamics and the Neurodegenerative Triad of Alzheimer’s Disease: The Hidden Connection. J. Neurochem. 2017, 143, 409–417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Mortal, S. Microtubule Dynamics in Cytoskeleton, Neurodegenerative and Psychiatric Disease. STEMedicine 2021, 2, e81. [Google Scholar] [CrossRef]
  63. Binet, S.; Meininger, V. Modifications of Microtubule Proteins in ALS Nerve Precede Detectable Histologic and Ultrastructural Changes. Neurology 1988, 38, 1596. [Google Scholar] [CrossRef] [PubMed]
  64. Fanara, P.; Banerjee, J.; Hueck, R.V.; Harper, M.R.; Awada, M.; Turner, H.; Husted, K.H.; Brandt, R.; Hellerstein, M.K. Stabilization of Hyperdynamic Microtubules Is Neuroprotective in Amyotrophic Lateral Sclerosis. J. Biol. Chem. 2007, 282, 23465–23472. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Ren, Y.; Jiang, H.; Yang, F.; Nakaso, K.; Feng, J. Parkin Protects Dopaminergic Neurons against Microtubule-Depolymerizing Toxins by Attenuating Microtubule-Associated Protein Kinase Activation. J. Biol. Chem. 2009, 284, 4009–4017. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Trushina, E.; Heldebrant, M.P.; Perez-Terzic, C.M.; Bortolon, R.; Kovtun, I.V.; Badger, J.D.; Terzic, A.; Estevez, A.; Windebank, A.J.; Dyer, R.B.; et al. Microtubule Destabilization and Nuclear Entry Are Sequential Steps Leading to Toxicity in Huntington’s Disease. Proc. Natl. Acad. Sci. USA 2003, 100, 12171–12176. [Google Scholar] [CrossRef] [Green Version]
  67. Hasegawa, M.; Smith, M.J.; Goedert, M. Tau Proteins with FTDP-17 Mutations Have a Reduced Ability to Promote Microtubule Assembly. FEBS Lett. 1998, 437, 207–210. [Google Scholar] [CrossRef] [Green Version]
  68. Alonso, A.D.C.; Grundke-Iqbal, I.; Iqbal, K. Alzheimer’s Disease Hyperphosphorylated Tau Sequesters Normal Tau into Tangles of Filaments and Disassembles Microtubules. Nat. Med. 1996, 2, 783–787. [Google Scholar] [CrossRef]
  69. Kaufmann, W.E.; Macdonald, S.M.; Altamura, C.R. Dendritic Cytoskeletal Protein Expression in Mental Retardation: An Immunohistochemical Study of the Neocortex in Rett Syndrome. Cereb. Cortex 2000, 10, 992–1004. [Google Scholar] [CrossRef] [Green Version]
  70. Gozes, I. Microtubules (Tau) as an Emerging Therapeutic Target: NAP (Davunetide). Curr. Pharm. Des. 2011, 17, 3413–3417. [Google Scholar] [CrossRef]
  71. Shelton, M.A.; Newman, J.T.; Gu, H.; Sampson, A.R.; Fish, K.N.; MacDonald, M.L.; Moyer, C.E.; DiBitetto, J.V.; Dorph-Petersen, K.-A.; Penzes, P.; et al. Loss of Microtubule-Associated Protein 2 Immunoreactivity Linked to Dendritic Spine Loss in Schizophrenia. Biol. Psychiatry 2015, 78, 374–385. [Google Scholar] [CrossRef] [Green Version]
  72. Drago, A.; Crisafulli, C.; Sidoti, A.; Calabrò, M.; Serretti, A. The Microtubule-Associated Molecular Pathways May Be Genetically Disrupted in Patients with Bipolar Disorder. Insights from the Molecular Cascades. J. Affect. Dis. 2016, 190, 429–438. [Google Scholar] [CrossRef] [PubMed]
  73. Cavaletti, G.; Marmiroli, P. Chemotherapy-Induced Peripheral Neurotoxicity. Nat. Rev. Neurol. 2010, 6, 657–666. [Google Scholar] [CrossRef]
  74. Mandilaras, V.; Wan-Chow-Wah, D.; Monette, J.; Gaba, F.; Monette, M.; Alfonso, L. The Impact of Cancer Therapy on Cognition in the Elderly. Front. Pharmacol. 2013, 4, 48. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Wu, Q.-F.; Yang, L.; Li, S.; Wang, Q.; Yuan, X.-B.; Gao, X.; Bao, L.; Zhang, X. Fibroblast Growth Factor 13 Is a Microtubule-Stabilizing Protein Regulating Neuronal Polarization and Migration. Cell 2012, 149, 1549–1564. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Cassimeris, L.; Spittle, C. Regulation of Microtubule-Associated Proteins. Int. Rev. Cytol. 2001, 210, 163–226. [Google Scholar] [PubMed]
  77. Fournet, V.; de Lavilléon, G.; Schweitzer, A.; Giros, B.; Andrieux, A.; Martres, M.-P. Both Chronic Treatments by Epothilone D and Fluoxetine Increase the Short-Term Memory and Differentially Alter the Mood Status of STOP/MAP6 KO Mice. J. Neurochem. 2012, 123, 982–996. [Google Scholar] [CrossRef] [Green Version]
  78. Neve, R.L.; Harris, P.; Kosik, K.S.; Kurnit, D.M.; Donlon, T.A. Identification of CDNA Clones for the Human Microtubule-Associated Protein Tau and Chromosomal Localization of the Genes for Tau and Microtubule-Associated Protein 2. Mol. Brain Res. 1986, 1, 271–280. [Google Scholar] [CrossRef]
  79. Guo, T.; Noble, W.; Hanger, D.P. Roles of Tau Protein in Health and Disease. Acta Neuropathol. 2017, 133, 665–704. [Google Scholar] [CrossRef] [Green Version]
  80. Lee, G.; Neve, R.L.; Kosik, K.S. The Microtubule Binding Domain of Tau Protein. Neuron 1989, 2, 1615–1624. [Google Scholar] [CrossRef]
  81. Amos, L.A. Microtubule Structure and Its Stabilisation. Org. Biomol. Chem. 2004, 2, 2153. [Google Scholar] [CrossRef] [PubMed]
  82. Drechsel, D.N.; Hyman, A.A.; Cobb, M.H.; Kirschner, M.W. Modulation of the Dynamic Instability of Tubulin Assembly by the Microtubule-Associated Protein Tau. Mol. Biol. Cell 1992, 3, 1141–1154. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Papasozomenos, S.C.; Binder, L.I. Phosphorylation Determines Two Distinct Species of Tau in the Central Nervous System. Cell Motil. Cytoskelet. 1987, 8, 210–226. [Google Scholar] [CrossRef] [PubMed]
  84. Ittner, L.M.; Ke, Y.D.; Delerue, F.; Bi, M.; Gladbach, A.; van Eersel, J.; Wölfing, H.; Chieng, B.C.; Christie, M.J.; Napier, I.A.; et al. Dendritic Function of Tau Mediates Amyloid-β Toxicity in Alzheimer’s Disease Mouse Models. Cell 2010, 142, 387–397. [Google Scholar] [CrossRef] [Green Version]
  85. Martin, L.; Latypova, X.; Wilson, C.M.; Magnaudeix, A.; Perrin, M.-L.; Terro, F. Tau Protein Phosphatases in Alzheimer’s Disease: The Leading Role of PP2A. Ageing Res. Rev. 2013, 12, 39–49. [Google Scholar] [CrossRef]
  86. Martin, L.; Latypova, X.; Wilson, C.M.; Magnaudeix, A.; Perrin, M.-L.; Yardin, C.; Terro, F. Tau Protein Kinases: Involvement in Alzheimer’s Disease. Ageing Res. Rev. 2013, 12, 289–309. [Google Scholar] [CrossRef]
  87. 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]
  88. Alonso, A.C.; Zaidi, T.; Grundke-Iqbal, I.; Iqbal, K. Role of Abnormally Phosphorylated Tau in the Breakdown of Microtubules in Alzheimer Disease. Proc. Natl. Acad. Sci. USA 1994, 91, 5562–5566. [Google Scholar] [CrossRef] [Green Version]
  89. Merrick, S.E.; Trojanowski, J.Q.; Lee, V.M.-Y. Selective Destruction of Stable Microtubules and Axons by Inhibitors of Protein Serine/Threonine Phosphatases in Cultured Human Neurons (NT2N Cells). J. Neurosci. 1997, 17, 5726–5737. [Google Scholar] [CrossRef]
  90. Zhang, B.; Maiti, A.; Shively, S.; Lakhani, F.; McDonald-Jones, G.; Bruce, J.; Lee, E.B.; Xie, S.X.; Joyce, S.; Li, C.; et al. Microtubule-Binding Drugs Offset Tau Sequestration by Stabilizing Microtubules and Reversing Fast Axonal Transport Deficits in a Tauopathy Model. Proc. Natl. Acad. Sci. USA 2004, 102, 227–231. [Google Scholar] [CrossRef] [Green Version]
  91. Alonso, A.d.C.; Grundke-Iqbal, I.; Barra, H.S.; Iqbal, K. Abnormal Phosphorylation of Tau and the Mechanism of Alzheimer Neurofibrillary Degeneration: Sequestration of Microtubule-Associated Proteins 1 and 2 and the Disassembly of Microtubules by the Abnormal Tau. Proc. Natl. Acad. Sci. USA 1997, 94, 298–303. [Google Scholar] [CrossRef] [Green Version]
  92. Hoover, B.R.; Reed, M.N.; Su, J.; Penrod, R.D.; Kotilinek, L.A.; Grant, M.K.; Pitstick, R.; Carlson, G.A.; Lanier, L.M.; Yuan, L.L.; et al. Tau Mislocalization to Dendritic Spines Mediates Synaptic Dysfunction Independently of Neurodegeneration. Neuron 2010, 68, 1067–1081. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Tackenberg, C.; Brandt, R. Divergent Pathways Mediate Spine Alterations and Cell Death Induced by Amyloid-, Wild-Type Tau, and R406W Tau. J. Neurosci. 2009, 29, 14439–14450. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Golovyashkina, N.; Penazzi, L.; Ballatore, C.; Smith, A.B.; Bakota, L.; Brandt, R. Region-Specific Dendritic Simplification Induced by Aβ, Mediated by Tau via Dysregulation of Microtubule Dynamics: A Mechanistic Distinct Event from Other Neurodegenerative Processes. Mol. Neurodegener. 2015, 10, 60. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Arendt, T.; Bullmann, T. Neuronal Plasticity in Hibernation and the Proposed Role of the Microtubule-Associated Protein Tau as a “Master Switch” Regulating Synaptic Gain in Neuronal Networks. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2013, 305, R478–R489. [Google Scholar] [CrossRef] [Green Version]
  96. Varidaki, A.; Hong, Y.; Coffey, E.T. Repositioning Microtubule Stabilizing Drugs for Brain Disorders. Front. Cell. Neurosci. 2018, 12, 226. [Google Scholar] [CrossRef] [PubMed]
  97. Goodin, S.; Kane, M.P.; Rubin, E.H. Epothilones: Mechanism of Action and Biologic Activity. J. Clin. Oncol. 2004, 22, 2015–2025. [Google Scholar] [CrossRef]
  98. Kolman, A. Epothilone D (Kosan/Roche). Curr. Opin. Investig. Drugs 2004, 5, 657–667. [Google Scholar]
  99. Michaelis, M.L. Ongoing In Vivo Studies with Cytoskeletal Drugs in Tau Transgenic Mice. Curr. Alzheimer Res. 2006, 3, 215–219. [Google Scholar] [CrossRef]
  100. Michaelis, M.L.; Ansar, S.; Chen, Y.; Reiff, E.R.; Seyb, K.I.; Himes, R.H.; Audus, K.L.; Georg, G.I. β-Amyloid-Induced Neurodegeneration and Protection by Structurally Diverse Microtubule-Stabilizing Agents. J. Pharmacol. Exp. Ther. 2005, 312, 659–668. [Google Scholar] [CrossRef]
  101. Shemesh, O.A.; Spira, M.E. Rescue of Neurons from Undergoing Hallmark Tau-Induced Alzheimer’s Disease Cell Pathologies by the Antimitotic Drug Paclitaxel. Neurobiol. Dis. 2011, 43, 163–175. [Google Scholar] [CrossRef] [PubMed]
  102. Silva, D.F.F.; Esteves, A.R.; Arduino, D.M.; Oliveira, C.R.; Cardoso, S.M. Amyloid-β-Induced Mitochondrial Dysfunction Impairs the Autophagic Lysosomal Pathway in a Tubulin Dependent Pathway. J. Alzheimers Dis. 2011, 26, 565–581. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Lee, J.J.; Swain, S.M. Peripheral Neuropathy Induced by Microtubule-Stabilizing Agents. J. Clin. Oncol. 2006, 24, 1633–1642. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Carlson, K.; Ocean, A.J. Peripheral Neuropathy with Microtubule-Targeting Agents: Occurrence and Management Approach. Clin. Breast Cancer 2011, 11, 73–81. [Google Scholar] [CrossRef] [Green Version]
  105. Chiorazzi, A.; Nicolini, G.; Canta, A.; Oggioni, N.; Rigolio, R.; Cossa, G.; Lombardi, R.; Roglio, I.; Cervellini, I.; Lauria, G.; et al. Experimental Epothilone B Neurotoxicity: Results of in Vitro and in Vivo Studies. Neurobiol. Dis. 2009, 35, 270–277. [Google Scholar] [CrossRef]
  106. LaPointe, N.E.; Morfini, G.; Brady, S.T.; Feinstein, S.C.; Wilson, L.; Jordan, M.A. Effects of Eribulin, Vincristine, Paclitaxel and Ixabepilone on Fast Axonal Transport and Kinesin-1 Driven Microtubule Gliding: Implications for Chemotherapy-Induced Peripheral Neuropathy. NeuroToxicology 2013, 37, 231–239. [Google Scholar] [CrossRef] [Green Version]
  107. Ballatore, C.; Brunden, K.R.; Trojanowski, J.Q.; Lee, V.M.Y.; Smith, A.B. Non-Naturally Occurring Small Molecule Microtubule-Stabilizing Agents: A Potential Tactic for CNS-Directed Therapies. ACS Chem. Neurosci. 2017, 8, 5–7. [Google Scholar] [CrossRef] [Green Version]
  108. Nogales, E.; Grayer Wolf, S.; Khan, I.A.; Ludueña, R.F.; Downing, K.H. Structure of Tubulin at 6.5 Å and Location of the Taxol-Binding Site. Nature 1995, 375, 424–427. [Google Scholar]
  109. Nogales, E.; Wolf, S.G.; Downing, K.H. Erratum: Structure of the Aβ Tubulin Dimer by Electron Crystallography. Nature 1998, 393, 191. [Google Scholar] [CrossRef] [Green Version]
  110. Amos, L.A.; Löwe, J. How Taxol® Stabilises Microtubule Structure. Chem. Biol. 1999, 6, R65–R69. [Google Scholar] [CrossRef] [Green Version]
  111. Prota, A.E.; Bargsten, K.; Zurwerra, D.; Field, J.J.; Díaz, J.F.; Altmann, K.-H.; Steinmetz, M.O. Molecular Mechanism of Action of Microtubule-Stabilizing Anticancer Agents. Science 2013, 339, 587–590. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Brunden, K.R.; Zhang, B.; Carroll, J.; Yao, Y.; Potuzak, J.S.; Hogan, A.M.L.; Iba, M.; James, M.J.; Xie, S.X.; Ballatore, C.; et al. Epothilone D Improves Microtubule Density, Axonal Integrity, and Cognition in a Transgenic Mouse Model of Tauopathy. J. Neurosci. 2010, 30, 13861–13866. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Hellal, F.; Hurtado, A.; Ruschel, J.; Flynn, K.C.; Laskowski, C.J.; Umlauf, M.; Kapitein, L.C.; Strikis, D.; Lemmon, V.; Bixby, J.; et al. Microtubule Stabilization Reduces Scarring and Causes Axon Regeneration After Spinal Cord Injury. Science 2011, 331, 928–931. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Sengottuvel, V.; Leibinger, M.; Pfreimer, M.; Andreadaki, A.; Fischer, D. Taxol Facilitates Axon Regeneration in the Mature CNS. J. Neurosci. 2011, 31, 2688–2699. [Google Scholar] [CrossRef]
  115. Baas, P.W.; Ahmad, F.J. Beyond Taxol: Microtubule-Based Treatment of Disease and Injury of the Nervous System. Brain 2013, 136, 2937–2951. [Google Scholar] [CrossRef] [Green Version]
  116. Flatters, S.J.L.; Bennett, G.J. Studies of Peripheral Sensory Nerves in Paclitaxel-Induced Painful Peripheral Neuropathy: Evidence for Mitochondrial Dysfunction. Pain 2006, 122, 245–257. [Google Scholar] [CrossRef] [Green Version]
  117. Markman, M.; Mekhail, T.M. Paclitaxel in Cancer Therapy. Exp. Opin. Pharmacother. 2002, 3, 755–766. [Google Scholar] [CrossRef]
  118. Alloatti, G.; Penna, C.; Gallo, M.P.; Levi, R.C.; Bombardelli, E.; Appendino, G. Differential Effects of Paclitaxel and Derivatives on Guinea Pig Isolated Heart and Papillary Muscle. J. Pharmacol. Exp. Ther. 1998, 284, 561. [Google Scholar]
  119. Gallo, J.M.; Li, S.; Guo, P.; Reed, K.; Ma, J. The Effect of P-Glycoprotein on Paclitaxel Brain and Brain Tumor Distribution in Mice. Cancer Res. 2003, 63, 5114. [Google Scholar]
  120. Kemper, E.M.; van Zandbergen, A.E.; Cleypool, C.; Mos, H.A.; Boogerd, W.; Beijnen, J.H.; van Tellingen, O. Increased Penetration of Paclitaxel into the Brain by Inhibition of P-Glycoprotein. Clin. Cancer Res. 2003, 9, 2849. [Google Scholar]
  121. Furukawa, K.; Mattson, M.P. Taxol Stabilizes [Ca2+]i and Protects Hippocampal Neurons against Excitotoxicity. Brain Res. 1995, 689, 141–146. [Google Scholar] [CrossRef]
  122. Wolf, S.; Barton, D.; Kottschade, L.; Grothey, A.; Loprinzi, C. Chemotherapy-Induced Peripheral Neuropathy: Prevention and Treatment Strategies. Eur. J. Cancer 2008, 44, 1507–1515. [Google Scholar] [CrossRef]
  123. Postma, T.J.; Hoekman, K.; van Riel, J.M.G.H.; Heimans, J.J.; Vermorken, J.B. Peripheral Neuropathy Due to Biweekly Paclitaxel, Epirubicin and Cisplatin in Patients with Advanced Ovarian Cancer. J. Neurooncol. 1999, 45, 241–246. [Google Scholar] [CrossRef]
  124. Ahles, T.A.; Saykin, A.J.; Furstenberg, C.T.; Cole, B.; Mott, L.A.; Skalla, K.; Whedon, M.B.; Bivens, S.; Mitchell, T.; Greenberg, E.R.; et al. Neuropsychologic Impact of Standard-Dose Systemic Chemotherapy in Long-Term Survivors of Breast Cancer and Lymphoma. J. Clin. Oncol. 2002, 20, 485–493. [Google Scholar] [CrossRef] [PubMed]
  125. Wefel, J.S.; Schagen, S.B. Chemotherapy-Related Cognitive Dysfunction. Curr. Neurol. Neurosci. Rep. 2012, 12, 267–275. [Google Scholar] [CrossRef] [PubMed]
  126. Ferris, C.F.; Nodine, S.; Pottala, T.; Cai, X.; Knox, T.M.; Fofana, F.H.; Kim, S.; Kulkarni, P.; Crystal, J.D.; Hohmann, A.G. Alterations in Brain Neurocircuitry Following Treatment with the Chemotherapeutic Agent Paclitaxel in Rats. Neurobiol. Pain 2019, 6, 100034. [Google Scholar] [CrossRef]
  127. Otová, B.; Václavíková, R.; Danielová, V.; Holubová, J.; Ehrlichová, M.; Horský, S.; Souček, P.; Šimek, P.; Gut, I. Effects of Paclitaxel, Docetaxel and Their Combinations on Subcutaneous Lymphomas in Inbred Sprague–Dawley/Cub Rats. Eur. J. Pharm. Sci. 2006, 29, 442–450. [Google Scholar] [CrossRef]
  128. Persohn, E.; Canta, A.; Schoepfer, S.; Traebert, M.; Mueller, L.; Gilardini, A.; Galbiati, S.; Nicolini, G.; Scuteri, A.; Lanzani, F.; et al. Morphological and Morphometric Analysis of Paclitaxel and Docetaxel-Induced Peripheral Neuropathy in Rats. Eur. J. Cancer 2005, 41, 1460–1466. [Google Scholar] [CrossRef]
  129. Park, S.R.; Kim, H.K.; Kim, C.G.; Choi, I.J.; Lee, J.S.; Lee, J.H.; Ryu, K.W.; Kim, Y.-W.; Bae, J.-M.; Kim, N.K. Phase I/II Study of S-1 Combined with Weekly Docetaxel in Patients with Metastatic Gastric Carcinoma. Br. J. Cancer 2008, 98, 1305–1311. [Google Scholar] [CrossRef]
  130. Ziske, C.G.; Schöttker, B.; Gorschlüter, M.; Mey, U.; Kleinschmidt, R.; Schlegel, U.; Sauerbruch, T.; Schmidt-Wolf, I.G.H. Acute Transient Encephalopathy after Paclitaxel Infusion: Report of Three Cases. Ann. Oncol. 2002, 13, 629–631. [Google Scholar] [CrossRef]
  131. Mercado-Gómez, O.; Ferrera, P.; Arias, C. Histopathologic Changes Induced by the Microtubule-Stabilizing Agent Taxol in the Rat Hippocampus In Vivo. J. Neurosci. Res. 2004, 78, 553–562. [Google Scholar] [CrossRef] [PubMed]
  132. Bollag, D.M.; McQueney, P.A.; Zhu, J.; Hensens, O.; Koupal, L.; Liesch, J.; Goetz, M.; Lazarides, E.; Woods, C.M. Epothilones, a New Class of Microtubule-Stabilizing Agents with a Taxol-like Mechanism of Action. Cancer Res. 1995, 55, 2325–2333. [Google Scholar] [PubMed]
  133. Giannakakou, P.; Gussio, R.; Nogales, E.; Downing, K.H.; Zaharevitz, D.; Bollbuck, B.; Poy, G.; Sackett, D.; Nicolaou, K.C.; Fojo, T. A Common Pharmacophore for Epothilone and Taxanes: Molecular Basis for Drug Resistance Conferred by Tubulin Mutations in Human Cancer Cells. Proc. Natl. Acad. Sci. USA 2000, 97, 2904–2909. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Nettles, J.H.; Li, H.; Cornett, B.; Krahn, J.M.; Snyder, J.P.; Downing, K.H. The Binding Mode of Epothilone A on α,ß-Tubulin by Electron Crystallography. Science 2004, 305, 866–869. [Google Scholar] [CrossRef]
  135. Andrieux, A.; Salin, P.; Schweitzer, A.; Bégou, M.; Pachoud, B.; Brun, P.; Gory-Fauré, S.; Kujala, P.; Suaud-Chagny, M.-F.; Höfle, G.; et al. Microtubule Stabilizer Ameliorates Synaptic Function and Behavior in a Mouse Model for Schizophrenia. Biol. Psychiatry 2006, 60, 1224–1230. [Google Scholar] [CrossRef] [Green Version]
  136. Brizuela, M.; Blizzard, C.A.; Chuckowree, J.A.; Dawkins, E.; Gasperini, R.J.; Young, K.M.; Dickson, T.C. The Microtubule-Stabilizing Drug Epothilone D Increases Axonal Sprouting Following Transection Injury in Vitro. Mol. Cell. Neurosci. 2015, 66, 129–140. [Google Scholar] [CrossRef]
  137. Sandner, B.; Puttagunta, R.; Motsch, M.; Bradke, F.; Ruschel, J.; Blesch, A.; Weidner, N. Systemic Epothilone D Improves Hindlimb Function after Spinal Cord Contusion Injury in Rats. Exp. Neurol. 2018, 306, 250–259. [Google Scholar] [CrossRef]
  138. Cartelli, D.; Casagrande, F.; Busceti, C.L.; Bucci, D.; Molinaro, G.; Traficante, A.; Passarella, D.; Giavini, E.; Pezzoli, G.; Battaglia, G.; et al. Microtubule Alterations Occur Early in Experimental Parkinsonism and The Microtubule Stabilizer Epothilone D Is Neuroprotective. Sci. Rep. 2013, 3, 1837. [Google Scholar] [CrossRef] [Green Version]
  139. Zhang, B.; Carroll, J.; Trojanowski, J.Q.; Yao, Y.; Iba, M.; Potuzak, J.S.; Hogan, A.M.L.; Xie, S.X.; Ballatore, C.; Smith, A.B.; et al. The Microtubule-Stabilizing Agent, Epothilone D, Reduces Axonal Dysfunction, Neurotoxicity, Cognitive Deficits, and Alzheimer-like Pathology in an Interventional Study with Aged Tau Transgenic Mice. J. Neurosci. 2012, 32, 3601–3611. [Google Scholar] [CrossRef]
  140. Chuckowree, J.A.; Zhu, Z.; Brizuela, M.; Lee, K.M.; Blizzard, C.A.; Dickson, T.C. The Microtubule-Modulating Drug Epothilone D Alters Dendritic Spine Morphology in a Mouse Model of Mild Traumatic Brain Injury. Front. Cell. Neurosci. 2018, 12, 223. [Google Scholar] [CrossRef] [Green Version]
  141. Ruschel, J.; Hellal, F.; Flynn, K.C.; Dupraz, S.; Elliott, D.A.; Tedeschi, A.; Bates, M.; Sliwinski, C.; Brook, G.; Dobrindt, K.; et al. Systemic Administration of Epothilone B Promotes Axon Regeneration after Spinal Cord Injury. Science 2015, 348, 347–352. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Jang, E.-H.; Sim, A.; Im, S.-K.; Hur, E.-M. Effects of Microtubule Stabilization by Epothilone B Depend on the Type and Age of Neurons. Neural Plast. 2016, 2016, 5056418. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Yang, Y.; Zhang, X.; Ge, H.; Liu, W.; Sun, E.; Ma, Y.; Zhao, H.; Li, R.; Chen, W.; Yuan, J.; et al. Epothilone B Benefits Nigrostriatal Pathway Recovery by Promoting Microtubule Stabilization After Intracerebral Hemorrhage. J. Am. Heart Assoc. 2018, 7, e007626. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Mao, L.; Gao, W.; Chen, S.; Song, Y.; Song, C.; Zhou, Z.; Zhao, H.; Zhou, K.; Wang, W.; Zhu, K.; et al. Epothilone B Impairs Functional Recovery after Spinal Cord Injury by Increasing Secretion of Macrophage Colony-Stimulating Factor. Cell Death Dis. 2017, 8, e3162. [Google Scholar] [CrossRef]
  145. Clark, J.A.; Blizzard, C.A.; Breslin, M.C.; Yeaman, E.J.; Lee, K.M.; Chuckowree, J.A.; Dickson, T.C. Epothilone D Accelerates Disease Progression in the SOD1 G93A Mouse Model of Amyotrophic Lateral Sclerosis. Neuropathol. Appl. Neurobiol. 2018, 44, 590–605. [Google Scholar] [CrossRef]
  146. Clark, J.; Zhu, Z.; Chuckowree, J.; Dickson, T.; Blizzard, C. Efficacy of epothilones in central nervous system trauma treatment: What has age got to do with it? Neural Regen. Res. 2021, 16, 618–620. [Google Scholar]
  147. Zhu, Z.; Chuckowree, J.A.; Musgrove, R.; Dickson, T.C.; Blizzard, C.A. The pathologic outcomes and efficacy of epothilone treatment following traumatic brain injury is determined by age. Neurobiol. Aging 2020, 93, 85–96. [Google Scholar] [CrossRef]
  148. Dollé, J.; Jayea, A.; Andersonb, S.A.; Ahmadzadehc, H.; Shenoyc, V.B.; Smith, D.H. Newfound sex differences in axonal structure underlie differential outcomes from in vitro traumatic axonal injury. Exp. Neurol. 2018, 300, 121–134. [Google Scholar] [CrossRef]
  149. Sahenk, Z.; Brady, S.T.; Mendell, J.R. Studies on the Pathogenesis of Vincristine-Induced Neuropathy. Muscle Nerve 1987, 10, 80–84. [Google Scholar] [CrossRef]
  150. Dumontet, C.; Jordan, M.A. Microtubule-Binding Agents: A Dynamic Field of Cancer Therapeutics. Nat. Rev. Drug Discov. 2010, 9, 790–803. [Google Scholar] [CrossRef] [Green Version]
  151. Escuin, D.; Kline, E.R.; Giannakakou, P. Both Microtubule-Stabilizing and Microtubule-Destabilizing Drugs Inhibit Hypoxia-Inducible Factor-1α Accumulation and Activity by Disrupting Microtubule Function. Cancer Res. 2005, 65, 9021–9028. [Google Scholar] [CrossRef] [Green Version]
  152. Jordan, M. Mechanism of Action of Antitumor Drugs That Interact with Microtubules and Tubulin. Curr. Med. Chem. Anticancer Agents 2012, 2, 1–17. [Google Scholar] [CrossRef] [PubMed]
  153. You, Z.; Zhang, S.; Shen, S.; Yang, J.; Ding, W.; Yang, L.; Lim, G.; Doheny, J.T.; Tate, S.; Chen, L.; et al. Cognitive Impairment in a Rat Model of Neuropathic Pain: Role of Hippocampal Microtubule Stability. Pain 2018, 159, 1518–1528. [Google Scholar] [CrossRef] [PubMed]
  154. O’Connell, C.; O’Malley, A.; Regan, C.M. Transient, Learning-Induced Ultrastructural Change in Spatially-Clustered Dentate Granule Cells of the Adult Rat Hippocampus. Neuroscience 1996, 76, 55–62. [Google Scholar] [CrossRef]
  155. Nelson, T.J.; Backlund, P.S.; Alkon, D.L. Hippocampal Protein-Protein Interactions in Spatial Memory. Hippocampus 2004, 14, 46–57. [Google Scholar] [CrossRef] [PubMed]
  156. Priel, A.; Tuszynski, J.A.; Woolf, N.J. Neural Cytoskeleton Capabilities for Learning and Memory. J. Biol. Phys. 2010, 36, 3. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Cavallaro, S.; D’Agata, V.; Manickam, P.; Dufour, F.; Alkon, D.L. Memory-Specific Temporal Profiles of Gene Expression in the Hippocampus. Proc. Natl. Acad. Sci. USA 2002, 99, 16279–16284. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Yamaguchi, S.; Fujii-Taira, I.; Murakami, A.; Hirose, N.; Aoki, N.; Izawa, E.-I.; Fujimoto, Y.; Takano, T.; Matsushima, T.; Homma, K.J. Up-Regulation of Microtubule-Associated Protein 2 Accompanying the Filial Imprinting of Domestic Chicks (Gallus Gallus Domesticus). Brain Res. Bull. 2008, 76, 282–288. [Google Scholar] [CrossRef]
  159. Woolf, N.J.; Young, S.L.; Johnson, G.V.W.; Fanselow, M.S. Pavlovian Conditioning Alters Cortical Microtubule-Associated Protein-2. NeuroReport 1994, 5, 1045–1048. [Google Scholar] [CrossRef]
  160. Woolf, N.J. A Structural Basis for Memory Storage in Mammals. Prog. Neurobiol. 1998, 55, 59–77. [Google Scholar] [CrossRef] [Green Version]
  161. Woolf, N.J.; Zinnerman, M.D.; Johnson, G.V.W. Hippocampal Microtubule-Associated Protein-2 Alterations with Contextual Memory. Brain Res. 1999, 821, 241–249. [Google Scholar] [CrossRef]
  162. Li, H.; Xu, W.; Wang, D.; Wang, L.; Fang, Q.; Wan, X.; Zhang, J.; Hu, Y.; Li, H.; Zhang, J.; et al. 4R Tau Modulates Cocaine-Associated Memory through Adult Dorsal Hippocampal Neurogenesis. J. Neurosci. 2021, 41, 6753–6774. [Google Scholar] [CrossRef]
  163. Goedert, M.; Jakes, R. Expression of separate isoforms of human tau protein: Correlation with the tau pattern in brain and effects on tubulin polymerization. EMBO J. 1990, 9, 4225–4230. [Google Scholar] [CrossRef]
  164. Mondragón-Rodríguez, S.; Trillaud-Doppia, E.; Dudilot, A.; Bourgeois, C.; Lauzon, M.; Leclerc, N.; Boehm, J. Interaction of Endogenous Tau Protein with Synaptic Proteins Is Regulated by N-Methyl-d-Aspartate Receptor-Dependent Tau Phosphorylation. J. Biol. Chem. 2012, 287, 32040–32053. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Liu, H.; Zhang, J.; Zheng, P.; Zhang, Y. Altered Expression of MAP-2, GAP-43, and Synaptophysin in the Hippocampus of Rats with Chronic Cerebral Hypoperfusion Correlates with Cognitive Impairment. Mol. Brain Res. 2005, 139, 169–177. [Google Scholar] [CrossRef] [PubMed]
  166. Bianchi, M.; Fone, K.F.C.; Azmi, N.; Heidbreder, C.A.; Hagan, J.J.; Marsden, C.A. Isolation Rearing Induces Recognition Memory Deficits Accompanied by Cytoskeletal Alterations in Rat Hippocampus. Eur. J. Neurosci. 2006, 24, 2894–2902. [Google Scholar] [CrossRef]
  167. Shimada, A.; Tsuzuki, M.; Keino, H.; Satoh, M.; Chiba, Y.; Saitoh, Y.; Hosokawa, M. Apical Vulnerability to Dendritic Retraction in Prefrontal Neurones of Ageing SAMP10 Mouse: A Model of Cerebral Degeneration. Neuropathol. Appl. Neurobiol. 2006, 32, 1–14. [Google Scholar] [CrossRef]
  168. Hu, X.; Viesselmann, C.; Nam, S.; Merriam, E.; Dent, E.W. Activity-Dependent Dynamic Microtubule Invasion of Dendritic Spines. J. Neurosci. 2008, 28, 13094–13105. [Google Scholar] [CrossRef]
  169. Pandey, K.; Sharma, S.K. Activity-Dependent Acetylation of Alpha Tubulin in the Hippocampus. J. Mol. Neurosci. 2011, 45, 1–4. [Google Scholar] [CrossRef]
  170. Wu, D.; Jin, Y.; Shapiro, T.M.; Hinduja, A.; Baas, P.W.; Tom, V.J. Chronic Neuronal Activation Increases Dynamic Microtubules to Enhance Functional Axon Regeneration after Dorsal Root Crush Injury. Nat. Commun. 2020, 11, 6131. [Google Scholar] [CrossRef]
  171. Tryba, A.K.; Peña, F.; Ramirez, J.-M. Stabilization of Bursting in Respiratory Pacemaker Neurons. J. Neurosci. 2003, 23, 3538–3546. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Peña, F.; Tapia, R. Seizures and Neurodegeneration Induced by 4-Aminopyridine in Rat Hippocampus in Vivo: Role of Glutamate- and GABA-Mediated Neurotransmission and of Ion Channels. Neuroscience 2000, 101, 547–561. [Google Scholar] [CrossRef]
  173. Peña, F.; Tapia, R. Relationships Among Seizures, Extracellular Amino Acid Changes, and Neurodegeneration Induced by 4-Aminopyridine in Rat Hippocampus: A Microdialysis and Electroencephalographic Study. J. Neurochem. 2008, 72, 2006–2014. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Alvarez, J.; Ramirez, B.U. Axonal Microtubules: Their Regulation by the Electrical Activity of the Nerve. Neurosci. Lett. 1979, 15, 19–22. [Google Scholar] [CrossRef]
  175. Halpain, S.; Greengard, P. Activation of NMDA Receptors Induces Rapid Dephosphorylation of the Cytoskeletal Protein MAP2. Neuron 1990, 5, 237–246. [Google Scholar] [CrossRef]
  176. Montoro, R.J.; Díaz-Nido, J.; Avila, J.; López-Barneo, J. N-Methyl-d-Aspartate Stimulates the Dephosphorylation of the Microtubule-Associated Protein 2 and Potentiates Excitatory Synaptic Pathways in the Rat Hippocampus. Neuroscience 1993, 54, 859–871. [Google Scholar] [CrossRef]
  177. Quinlan, E.M.; Halpain, S. Postsynaptic Mechanisms for Bidirectional Control of MAP2 Phosphorylation by Glutamate Receptors. Neuron 1996, 16, 357–368. [Google Scholar] [CrossRef] [Green Version]
  178. Philpot, B.D.; Lim, J.H.; Halpain, S.; Brunjes, P.C. Experience-Dependent Modifications in MAP2 Phosphorylation in Rat Olfactory Bulb. J. Neurosci. 1997, 17, 9596–9604. [Google Scholar] [CrossRef] [Green Version]
  179. Kobayashi, S.; Tanaka, T.; Soeda, Y.; Takashima, A. Enhanced Tau Protein Translation by Hyper-Excitation. Front. Aging Neurosci. 2019, 11, 322. [Google Scholar] [CrossRef] [Green Version]
  180. Schätzle, P.; Esteves da Silva, M.; Tas, R.P.; Katrukha, E.A.; Hu, H.Y.; Wierenga, C.J.; Kapitein, L.C.; Hoogenraad, C.C. Activity-Dependent Actin Remodeling at the Base of Dendritic Spines Promotes Microtubule Entry. Curr. Biol. 2018, 28, 2081–2093e6. [Google Scholar] [CrossRef] [Green Version]
  181. Merriam, E.B.; Lumbard, D.C.; Viesselmann, C.; Ballweg, J.; Stevenson, M.; Pietila, L.; Hu, X.; Dent, E.W. Dynamic Microtubules Promote Synaptic NMDA Receptor-Dependent Spine Enlargement. PLoS ONE 2011, 6, e27688. [Google Scholar] [CrossRef] [PubMed]
  182. Merriam, E.B.; Millette, M.; Lumbard, D.C.; Saengsawang, W.; Fothergill, T.; Hu, X.; Ferhat, L.; Dent, E.W. Synaptic Regulation of Microtubule Dynamics in Dendritic Spines by Calcium, F-Actin, and Drebrin. J. Neurosci. 2013, 33, 16471–16482. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Mitsuyama, F.; Niimi, G.; Kato, K.; Hirosawa, K.; Mikoshiba, K.; Okuya, M.; Karagiozov, K.; Kato, Y.; Kanno, T.; Sanoe, H.; et al. Redistribution of Microtubules in Dendrites of Hippocampal CA1 Neurons after Tetanic Stimulation during Long-Term Potentiation. Arch. Ital. Anat. Embriol. 2008, 113, 17–27. [Google Scholar]
  184. Kapitein, L.C.; Yau, K.W.; Gouveia, S.M.; van der Zwan, W.A.; Wulf, P.S.; Keijzer, N.; Demmers, J.; Jaworski, J.; Akhmanova, A.; Hoogenraad, C.C. NMDA Receptor Activation Suppresses Microtubule Growth and Spine Entry. J. Neurosci. 2011, 31, 8194–8209. [Google Scholar] [CrossRef] [PubMed]
  185. Weisenberg, R.C. Microtubule Formation in Vitro in Solutions Containing Low Calcium Concentrations. Science 1972, 177, 1104–1105. [Google Scholar] [CrossRef]
  186. Fuller, G.M.; Brinkley, B.R. Structure and Control of Assembly of Cytoplasmic Microtubules in Normal and Transformed Cells. J. Supramol. Struct. 1976, 5, 497–514. [Google Scholar] [CrossRef]
  187. Schliwa, M. The Role of Divalent Cations in the Regulation of Microtubule Assembly: In Vivo Studies on Microtubules of the Heliozoan Axopodium Using the Ionophore A23187. J. Cell Biol. 1976, 70, 527–540. [Google Scholar]
  188. Marcum, J.M.; Dedman, J.R.; Brinkley, B.R.; Means, A.R. Control of Microtubule Assembly-Disassembly by Calcium-Dependent Regulator Protein. Proc. Natl. Acad. Sci. USA 1978, 75, 3771–3775. [Google Scholar] [CrossRef] [Green Version]
  189. Schliwa, M.; Euteneuer, U.; Bulinski, J.C.; Izant, J.G. Calcium Lability of Cytoplasmic Microtubules and Its Modulation by Microtubule-Associated Proteins. Proc. Natl. Acad. Sci. USA 1981, 78, 1037–1041. [Google Scholar] [CrossRef] [Green Version]
  190. Lee, Y.C.; Wolff, J. Two Opposing Effects of Calmodulin on Microtubule Assembly Depend on the Presence of Microtubule-Associated Proteins. J. Biol. Chem. 1982, 257, 6306–6310. [Google Scholar] [CrossRef]
  191. Deery, W.J.; Means, A.R.; Brinkley, B.R. Calmodulin-Microtubule Association in Cultured Mammalian Cells. J. Cell Biol. 1984, 98, 904–910. [Google Scholar] [CrossRef] [Green Version]
  192. Adamec, E.; Mercken, M.; Beermann, M.L.; Didier, M.; Nixon, R.A. Acute Rise in the Concentration of Free Cytoplasmic Calcium Leads to Dephosphorylation of the Microtubule-Associated Protein Tau. Brain Res. 1997, 757, 93–101. [Google Scholar] [CrossRef]
  193. Maas, C.; Belgardt, D.; Lee, H.K.; Heisler, F.F.; Lappe-Siefke, C.; Magiera, M.M.; van Dijk, J.; Hausrat, T.J.; Janke, C.; Kneussel, M. Synaptic Activation Modifies Microtubules Underlying Transport of Postsynaptic Cargo. Proc. Natl. Acad. Sci. USA 2009, 106, 8731–8736. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Xu, X.; Hu, Y.; Xiong, Y.; Li, Z.; Wang, W.; Du, C.; Yang, Y.; Zhang, Y.; Xiao, F.; Wang, X. Association of Microtubule Dynamics with Chronic Epilepsy. Mol. Neurobiol. 2016, 53, 5013–5024. [Google Scholar] [CrossRef]
  195. Wu, J.W.; Hussaini, S.A.; Bastille, I.M.; Rodriguez, G.A.; Mrejeru, A.; Rilett, K.; Sanders, D.W.; Cook, C.; Fu, H.; Boonen, R.A.C.M.; et al. Neuronal Activity Enhances Tau Propagation and Tau Pathology In Vivo. Nat. Neurosci. 2016, 19, 1085–1092. [Google Scholar] [CrossRef] [PubMed]
  196. Yamada, K.; Holth, J.K.; Liao, F.; Stewart, F.R.; Mahan, T.E.; Jiang, H.; Cirrito, J.R.; Patel, T.K.; Hochgräfe, K.; Mandelkow, E.-M.; et al. Neuronal Activity Regulates Extracellular Tau In Vivo. J. Exp. Med. 2014, 211, 387–393. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Pooler, A.M.; Phillips, E.C.; Lau, D.H.W.; Noble, W.; Hanger, D.P. Physiological Release of Endogenous Tau Is Stimulated by Neuronal Activity. EMBO Rep. 2013, 14, 389–394. [Google Scholar] [CrossRef]
  198. Froehner, S.C. Regulation of Ion Channel Distribution at Synapses. Ann. Rev. Neurosci. 1993, 16, 347–368. [Google Scholar] [CrossRef]
  199. Casini, S.; Tan, H.L.; Demirayak, I.; Remme, C.A.; Amin, A.S.; Scicluna, B.P.; Chatyan, H.; Ruijter, J.M.; Bezzina, C.R.; van Ginneken, A.C.G.; et al. Tubulin Polymerization Modifies Cardiac Sodium Channel Expression and Gating. Cardiovasc. Res. 2010, 85, 691–700. [Google Scholar] [CrossRef] [Green Version]
  200. Goswami, C.; Islam, M.S. Transient Receptor Potential Channels: What Is Happening? Reflections in the Wake of the 2009 TRP Meeting, Karolinska Institutet, Stockholm. Channels 2010, 4, 124–135. [Google Scholar]
  201. Johnson, B.D.; Byerly, L. A Cytoskeletal Mechanism for Ca2+ Channel Metabolic Dependence and Inactivation by Intracellular Ca2+. Neuron 1993, 10, 797–804. [Google Scholar] [CrossRef]
  202. Johnson, B.D.; Byerly, L. Ca2+ Channel Ca2+-Dependent Inactivation in a Mammalian Central Neuron Involves the Cytoskeleton. Pflug. Arch. 1994, 429, 14–21. [Google Scholar] [CrossRef] [PubMed]
  203. Pascarel, C.; Brette, F.; Cazorla, O.; le Guennec, J.-Y. Effects on L-Type Calcium Current of Agents Interfering with the Cytoskeleton of Isolated Guinea-Pig Ventricular Myocytes. Exp. Physiol. 1999, 84, 1043–1050. [Google Scholar] [CrossRef] [PubMed]
  204. Malan, D. Microtubules Mobility Affects the Modulation of L-Type ICa by Muscarinic and β-Adrenergic Agonists in Guinea-Pig Cardiac Myocytes. J. Mol. Cell. Cardiol. 2003, 35, 195–206. [Google Scholar] [CrossRef]
  205. Li, Z.; Hall, A.M.; Kelinske, M.; Roberson, E.D. Seizure Resistance without Parkinsonism in Aged Mice after Tau Reduction. Neurobiol. Aging 2014, 35, 2617–2624. [Google Scholar] [CrossRef] [Green Version]
  206. DeVos, S.L.; Goncharoff, D.K.; Chen, G.; Kebodeaux, C.S.; Yamada, K.; Stewart, F.R.; Schuler, D.R.; Maloney, S.E.; Wozniak, D.F.; Rigo, F.; et al. Antisense Reduction of Tau in Adult Mice Protects against Seizures. J. Neurosci. 2013, 33, 12887–12897. [Google Scholar] [CrossRef] [Green Version]
  207. Gheyara, A.L.; Ponnusamy, R.; Djukic, B.; Craft, R.J.; Ho, K.; Guo, W.; Finucane, M.M.; Sanchez, P.E.; Mucke, L. Tau Reduction Prevents Disease in a Mouse Model of Dravet Syndrome. Ann. Neurol. 2014, 76, 443–456. [Google Scholar] [CrossRef] [Green Version]
  208. Holth, J.K.; Bomben, V.C.; Reed, J.G.; Inoue, T.; Younkin, L.; Younkin, S.G.; Pautler, R.G.; Botas, J.; Noebels, J.L. Tau Loss Attenuates Neuronal Network Hyperexcitability in Mouse and Drosophila Genetic Models of Epilepsy. J. Neurosci. 2013, 33, 1651–1659. [Google Scholar] [CrossRef]
  209. Cloyd, R.A.; Koren, J.; Abisambra, J.F.; Smith, B.N. Effects of Altered Tau Expression on Dentate Granule Cell Excitability in Mice. Exp. Neurol. 2021, 343, 113766. [Google Scholar] [CrossRef]
  210. Hall, A.M.; Throesch, B.T.; Buckingham, S.C.; Markwardt, S.J.; Peng, Y.; Wang, Q.; Hoffman, D.A.; Roberson, E.D. Tau-Dependent Kv4.2 Depletion and Dendritic Hyperexcitability in a Mouse Model of Alzheimer’s Disease. J. Neurosci. 2015, 35, 6221–6230. [Google Scholar]
  211. Crimins, J.L.; Rocher, A.B.; Luebke, J.I. Electrophysiological Changes Precede Morphological Changes to Frontal Cortical Pyramidal Neurons in the RTg4510 Mouse Model of Progressive Tauopathy. Acta Neuropathol. 2012, 124, 777–795. [Google Scholar] [CrossRef] [PubMed]
  212. Crimins, J.L.; Rocher, A.B.; Peters, A.; Shultz, P.; Lewis, J.; Luebke, J.I. Homeostatic Responses by Surviving Cortical Pyramidal Cells in Neurodegenerative Tauopathy. Acta Neuropathol. 2011, 122, 551–564. [Google Scholar] [CrossRef] [PubMed]
  213. Rocher, A.B.; Crimins, J.L.; Amatrudo, J.M.; Kinson, M.S.; Todd-Brown, M.A.; Lewis, J.; Luebke, J.I. Structural and Functional Changes in Tau Mutant Mice Neurons Are Not Linked to the Presence of NFTs. Exp. Neurol. 2010, 223, 385–393. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. García-Cabrero, A.M.; Guerrero-López, R.; Giráldez, B.G.; Llorens-Martín, M.; Ávila, J.; Serratosa, J.M.; Sánchez, M.P. Hyperexcitability and Epileptic Seizures in a Model of Frontotemporal Dementia. Neurobiol. Dis. 2013, 58, 200–208. [Google Scholar] [CrossRef] [PubMed]
  215. Liu, X.; Ou, S.; Yin, M.; Xu, T.; Wang, T.; Liu, Y.; Ding, X.; Yu, X.; Yuan, J.; Huang, H.; et al. N-Methyl-D-Aspartate Receptors Mediate Epilepsy-Induced Axonal Impairment and Tau Phosphorylation via Activating Glycogen Synthase Kinase-3β and Cyclin-Dependent Kinase 5. Discov. Med. 2017, 23, 221–234. [Google Scholar] [PubMed]
  216. Pandis, D.; Scarmeas, N. Seizures in Alzheimer Disease: Clinical and Epidemiological Data. Epilepsy Curr. 2012, 12, 184–187. [Google Scholar] [CrossRef] [PubMed]
  217. Vossel, K.A.; Beagle, A.J.; Rabinovici, G.D.; Shu, H.; Lee, S.E.; Naasan, G.; Hegde, M.; Cornes, S.B.; Henry, M.L.; Nelson, A.B.; et al. Seizures and Epileptiform Activity in the Early Stages of Alzheimer Disease. JAMA Neurol. 2013, 70, 1158–1166. [Google Scholar] [CrossRef]
  218. Vossel, K.A.; Ranasinghe, K.G.; Beagle, A.J.; Mizuiri, D.; Honma, S.M.; Dowling, A.F.; Darwish, S.M.; van Berlo, V.; Barnes, D.E.; Mantle, M.; et al. Incidence and Impact of Subclinical Epileptiform Activity in Alzheimer’s Disease. Ann. Neurol. 2016, 80, 858–870. [Google Scholar] [CrossRef]
  219. Peña-Ortega, F. Brain Arrhythmias Induced by Amyloid Beta and Inflammation: Involvement in Alzheimer’s Disease and Other Inflammation-Related Pathologies. Curr. Alzheimer Res. 2019, 16, 1108–1131. [Google Scholar] [CrossRef]
  220. Menkes-Caspi, N.; Yamin, H.G.; Kellner, V.; Spires-Jones, T.L.; Cohen, D.; Stern, E.A. Pathological Tau Disrupts Ongoing Network Activity. Neuron 2015, 85, 959–966. [Google Scholar] [CrossRef] [Green Version]
  221. Hatch, R.J.; Wei, Y.; Xia, D.; Götz, J. Hyperphosphorylated Tau Causes Reduced Hippocampal CA1 Excitability by Relocating the Axon Initial Segment. Acta Neuropathol. 2017, 133, 717–730. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Busche, M.A.; Wegmann, S.; Dujardin, S.; Commins, C.; Schiantarelli, J.; Klickstein, N.; Kamath, T.V.; Carlson, G.A.; Nelken, I.; Hyman, B.T. Tau Impairs Neural Circuits, Dominating Amyloid-β Effects, in Alzheimer Models In Vivo. Nat. Neurosci. 2019, 22, 57–64. [Google Scholar] [CrossRef] [PubMed]
  223. Kopach, O.; Esteras, N.; Wray, S.; Rusakov, D.A.; Abramov, A.Y. Maturation and Phenotype of Pathophysiological Neuronal Excitability of Human Cells in Tau-Related Dementia. J. Cell Sci. 2020, 133, jcs241687. [Google Scholar] [CrossRef] [PubMed]
  224. Zempel, H.; Dennissen, F.J.A.; Kumar, Y.; Luedtke, J.; Biernat, J.; Mandelkow, E.-M.; Mandelkow, E. Axodendritic Sorting and Pathological Missorting of Tau Are Isoform-Specific and Determined by Axon Initial Segment Architecture. J. Biol. Chem. 2017, 292, 12192–12207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Schappacher, K.A.; Xie, W.; Zhang, J.-M.; Baccei, M.L. Neonatal Vincristine Administration Modulates Intrinsic Neuronal Excitability in the Rat Dorsal Root Ganglion and Spinal Dorsal Horn during Adolescence. Pain 2019, 160, 645–657. [Google Scholar] [CrossRef]
  226. Sun, S.; Zhang, H.; Liu, J.; Popugaeva, E.; Xu, N.-J.; Feske, S.; White, C.L.; Bezprozvanny, I. Reduced Synaptic STIM2 Expression and Impaired Store-Operated Calcium Entry Cause Destabilization of Mature Spines in Mutant Presenilin Mice. Neuron 2014, 82, 79–93. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Tsushima, H.; Emanuele, M.; Polenghi, A.; Esposito, A.; Vassalli, M.; Barberis, A.; Difato, F.; Chieregatti, E. HDAC6 and RhoA Are Novel Players in Abeta-Driven Disruption of Neuronal Polarity. Nat. Commun. 2015, 6, 7781. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Li, X.; Kumar, Y.; Zempel, H.; Mandelkow, E.-M.; Biernat, J.; Mandelkow, E. Novel Diffusion Barrier for Axonal Retention of Tau in Neurons and Its Failure in Neurodegeneration. EMBO J. 2011, 30, 4825–4837. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  229. Wang, Q.; Yang, J.; Wang, H.; Shan, B.; Yin, C.; Yu, H.; Zhang, X.; Dong, Z.; Yu, Y.; Zhao, R.; et al. Fibroblast Growth Factor 13 Stabilizes Microtubules to Promote Na+ Channel Function in Nociceptive DRG Neurons and Modulates Inflammatory Pain. J. Adv. Res. 2021, 31, 97–111. [Google Scholar] [CrossRef]
  230. Akin, E.J.; Alsaloum, M.; Higerd, G.P.; Liu, S.; Zhao, P.; Dib-Hajj, F.B.; Waxman, S.G.; Dib-Hajj, S.D. Paclitaxel Increases Axonal Localization and Vesicular Trafficking of Nav1.7. Brain 2021, 144, 1727–1737. [Google Scholar] [CrossRef]
  231. Illán-Gala, I.; Díaz de Terán, F.J.; Alonso, P.; Aguilar-Amat, M.-J. Nonconvulsive Status Epilepticus Secondary to Paclitaxel Administration. Epilepsy Behav. Case Rep. 2015, 4, 20–22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  232. Carletti, F.; Sardo, P.; Gambino, G.; Liu, X.-A.; Ferraro, G.; Rizzo, V. Hippocampal Hyperexcitability Is Modulated by Microtubule-Active Agent: Evidence from In Vivo and In Vitro Epilepsy Models in the Rat. Front. Cell. Neurosci. 2016, 10, 29. [Google Scholar] [CrossRef] [Green Version]
  233. Sohn, P.D.; Huang, C.T.L.; Yan, R.; Fan, L.; Tracy, T.E.; Camargo, C.M.; Montgomery, K.M.; Arhar, T.; Mok, S.A.; Freilich, R.; et al. Pathogenic Tau Impairs Axon Initial Segment Plasticity and Excitability Homeostasis. Neuron 2019, 104, 458–470.e5. [Google Scholar] [CrossRef] [PubMed]
  234. Evans, M.D.; Sammons, R.P.; Lebron, S.; Dumitrescu, A.S.; Watkins, T.B.K.; Uebele, V.N.; Renger, J.J.; Grubb, M.S. Calcineurin Signaling Mediates Activity-Dependent Relocation of the Axon Initial Segment. J. Neurosci. 2013, 33, 6950–6963. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Evans, M.D.; Dumitrescu, A.S.; Kruijssen, D.L.H.; Taylor, S.E.; Grubb, M.S. Rapid Modulation of Axon Initial Segment Length Influences Repetitive Spike Firing. Cell Rep. 2015, 13, 1233–1245. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Kuba, H.; Oichi, Y.; Ohmori, H. Presynaptic Activity Regulates Na+ Channel Distribution at the Axon Initial Segment. Nature 2010, 465, 1075–1078. [Google Scholar] [CrossRef]
  237. Yamada, R.; Kuba, H. Structural and Functional Plasticity at the Axon Initial Segment. Front. Cell. Neurosci. 2016, 10, 250. [Google Scholar] [CrossRef] [Green Version]
  238. Ogawa, Y.; Rasband, M.N. The Functional Organization and Assembly of the Axon Initial Segment. Curr. Opin. Neurobiol. 2008, 18, 307–313. [Google Scholar] [CrossRef]
  239. Jones, S.L.; Korobova, F.; Svitkina, T. Axon Initial Segment Cytoskeleton Comprises a Multiprotein Submembranous Coat Containing Sparse Actin Filaments. J. Cell Biol. 2014, 205, 67–81. [Google Scholar] [CrossRef] [Green Version]
  240. Leterrier, C.; Potier, J.; Caillol, G.; Debarnot, C.; Rueda Boroni, F.; Dargent, B. Nanoscale Architecture of the Axon Initial Segment Reveals an Organized and Robust Scaffold. Cell Rep. 2015, 13, 2781–2793. [Google Scholar] [CrossRef] [Green Version]
  241. Rasband, M.N. The Axon Initial Segment and the Maintenance of Neuronal Polarity. Nat. Rev. Neurosci. 2010, 11, 552–562. [Google Scholar] [CrossRef] [PubMed]
  242. Palay, S.L.; Sotelo, C.; Peters, A.; Orkand, P.M. The Axon Hillock And The Axon Initial Segment. J. Cell Biol. 1968, 38, 193–201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  243. Winckler, B.; Forscher, P.; Mellman, I. A Diffusion Barrier Maintains Distribution of Membrane Proteins in Polarized Neurons. Nature 1999, 397, 698–701. [Google Scholar] [CrossRef] [PubMed]
  244. Song, A.; Wang, D.; Chen, G.; Li, Y.; Luo, J.; Duan, S.; Poo, M. A Selective Filter for Cytoplasmic Transport at the Axon Initial Segment. Cell 2009, 136, 1148–1160. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Nakada, C.; Ritchie, K.; Oba, Y.; Nakamura, M.; Hotta, Y.; Iino, R.; Kasai, R.S.; Yamaguchi, K.; Fujiwara, T.; Kusumi, A. Accumulation of Anchored Proteins Forms Membrane Diffusion Barriers during Neuronal Polarization. Nat. Cell Biol. 2003, 5, 626–632. [Google Scholar] [CrossRef]
  246. Jenkins, S.M.; Bennett, V. Ankyrin-G Coordinates Assembly of the Spectrin-Based Membrane Skeleton, Voltage-Gated Sodium Channels, and L1 CAMs at Purkinje Neuron Initial Segments. J. Cell Biol. 2001, 155, 739–746. [Google Scholar] [CrossRef]
  247. Yang, Y.; Ogawa, Y.; Hedstrom, K.L.; Rasband, M.N. ΒIV Spectrin Is Recruited to Axon Initial Segments and Nodes of Ranvier by AnkyrinG. J. Cell Biol. 2007, 176, 509–519. [Google Scholar] [CrossRef]
  248. Bender, K.J.; Trussell, L.O. Axon Initial Segment Ca2+ Channels Influence Action Potential Generation and Timing. Neuron 2009, 61, 259–271. [Google Scholar] [CrossRef] [Green Version]
  249. Naundorf, B.; Wolf, F.; Volgushev, M. Unique Features of Action Potential Initiation in Cortical Neurons. Nature 2006, 440, 1060–1063. [Google Scholar] [CrossRef]
  250. Grubb, M.S.; Burrone, J. Activity-Dependent Relocation of the Axon Initial Segment Fine-Tunes Neuronal Excitability. Nature 2010, 465, 1070–1074. [Google Scholar] [CrossRef] [Green Version]
  251. Chand, A.N.; Galliano, E.; Chesters, R.A.; Grubb, M.S. A Distinct Subtype of Dopaminergic Interneuron Displays Inverted Structural Plasticity at the Axon Initial Segment. J. Neurosci. 2015, 35, 1573–1590. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  252. Sun, X.; Wu, Y.; Gu, M.; Liu, Z.; Ma, Y.; Li, J.; Zhang, Y. Selective Filtering Defect at the Axon Initial Segment in Alzheimer’s Disease Mouse Models. Proc. Natl. Acad. Sci. USA 2014, 111, 14271–14276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  253. Sohn, P.D.; Tracy, T.E.; Son, H.I.; Zhou, Y.; Leite, R.E.P.; Miller, B.L.; Seeley, W.W.; Grinberg, L.T.; Gan, L. Acetylated Tau Destabilizes the Cytoskeleton in the Axon Initial Segment and Is Mislocalized to the Somatodendritic Compartment. Mol. Neurodegener. 2016, 11, 47. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  254. Ramirez, J.-M.; Tryba, A.K.; Peña, F. Pacemaker Neurons and Neuronal Networks: An Integrative View. Curr. Opin. Neurobiol. 2004, 14, 665–674. [Google Scholar] [CrossRef] [PubMed]
  255. Morales, M.; Colicos, M.A.; Goda, Y. Actin-Dependent Regulation of Neurotransmitter Release at Central Synapses. Neuron 2000, 27, 539–550. [Google Scholar] [CrossRef] [Green Version]
  256. Lepicard, S.; Franco, B.; de Bock, F.; Parmentier, M.-L. A Presynaptic Role of Microtubule-Associated Protein 1/Futsch in Drosophila: Regulation of Active Zone Number and Neurotransmitter Release. J. Neurosci. 2014, 34, 6759–6771. [Google Scholar] [CrossRef] [Green Version]
  257. Watanabe, S.; Rost, B.R.; Camacho-Pérez, M.; Davis, M.W.; Söhl-Kielczynski, B.; Rosenmund, C.; Jorgensen, E.M. Ultrafast Endocytosis at Mouse Hippocampal Synapses. Nature 2013, 504, 242–247. [Google Scholar] [CrossRef] [Green Version]
  258. Delvendahl, I.; Vyleta, N.P.; von Gersdorff, H.; Hallermann, S. Fast, Temperature-Sensitive and Clathrin-Independent Endocytosis at Central Synapses. Neuron 2016, 90, 492–498. [Google Scholar] [CrossRef] [Green Version]
  259. Wu, X.-S.; Lee, S.H.; Sheng, J.; Zhang, Z.; Zhao, W.-D.; Wang, D.; Jin, Y.; Charnay, P.; Ervasti, J.M.; Wu, L.-G. Actin Is Crucial for All Kinetically Distinguishable Forms of Endocytosis at Synapses. Neuron 2016, 92, 1020–1035. [Google Scholar] [CrossRef] [Green Version]
  260. Sakaba, T.; Neher, E. Involvement of Actin Polymerization in Vesicle Recruitment at the Calyx of Held Synapse. J. Neurosci. 2003, 23, 837–846. [Google Scholar] [CrossRef] [Green Version]
  261. Cole, J.C.; Villa, B.R.S.; Wilkinson, R.S. Disruption of Actin Impedes Transmitter Release in Snake Motor Terminals. J. Physiol. 2000, 525, 579–586. [Google Scholar] [CrossRef] [PubMed]
  262. Lipstein, N.; Sakaba, T.; Cooper, B.H.; Lin, K.-H.; Strenzke, N.; Ashery, U.; Rhee, J.-S.; Taschenberger, H.; Neher, E.; Brose, N. Dynamic Control of Synaptic Vesicle Replenishment and Short-Term Plasticity by Ca2+-Calmodulin-Munc13-1 Signaling. Neuron 2013, 79, 82–96. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  263. Hosoi, N.; Holt, M.; Sakaba, T. Calcium Dependence of Exo- and Endocytotic Coupling at a Glutamatergic Synapse. Neuron 2009, 63, 216–229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Lee, J.S.; Ho, W.-K.; Lee, S.-H. Actin-Dependent Rapid Recruitment of Reluctant Synaptic Vesicles into a Fast-Releasing Vesicle Pool. Proc. Natl. Acad. Sci. USA 2012, 109, E765–E774. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  265. Lee, S.; Jung, K.J.; Jung, H.S.; Chang, S. Dynamics of Multiple Trafficking Behaviors of Individual Synaptic Vesicles Revealed by Quantum-Dot Based Presynaptic Probe. PLoS ONE 2012, 7, e38045. [Google Scholar] [CrossRef]
  266. Hirokawa, N.; Niwa, S.; Tanaka, Y. Molecular Motors in Neurons: Transport Mechanisms and Roles in Brain Function, Development, and Disease. Neuron 2010, 68, 610–638. [Google Scholar] [CrossRef] [Green Version]
  267. Melkov, A.; Abdu, U. Regulation of Long-Distance Transport of Mitochondria along Microtubules. Cell. Mol. Life Sci. 2018, 75, 163–176. [Google Scholar] [CrossRef]
  268. Gray, G. Synaptic Vesicles and Microtubules in Frog Motor Endplates. Proc. R. Soc. Lond. Ser. B Biol. Sci. 1978, 203, 219–227. [Google Scholar]
  269. Gray, G. Neurotransmitter Release Mechanisms and Microtubules. Proc. R. Soc. Lond. Ser. B Biol. Sci. 1983, 218, 253–258. [Google Scholar]
  270. Hirokawa, N.; Sobue, K.; Kanda, K.; Harada, A.; Yorifuji, H. The Cytoskeletal Architecture of the Presynaptic Terminal and Molecular Structure of Synapsin 1. J. Cell Biol. 1989, 108, 111–126. [Google Scholar] [CrossRef] [Green Version]
  271. Hummel, T.; Krukkert, K.; Roos, J.; Davis, G.; Klämbt, C. Drosophila Futsch/22C10 Is a MAP1B-like Protein Required for Dendritic and Axonal Development. Neuron 2000, 26, 357–370. [Google Scholar] [CrossRef] [Green Version]
  272. Perkins, G.A.; Tjong, J.; Brown, J.M.; Poquiz, P.H.; Scott, R.T.; Kolson, D.R.; Ellisman, M.H.; Spirou, G.A. The Micro-Architecture of Mitochondria at Active Zones: Electron Tomography Reveals Novel Anchoring Scaffolds and Cristae Structured for High-Rate Metabolism. J. Neurosci. 2010, 30, 1015–1026. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  273. Guillaud, L.; Dimitrov, D.; Takahashi, T. Presynaptic Morphology and Vesicular Composition Determine Vesicle Dynamics in Mouse Central Synapses. eLife 2017, 6, e24845. [Google Scholar] [CrossRef] [PubMed]
  274. Guedes-Dias, P.; Nirschl, J.J.; Abreu, N.; Tokito, M.K.; Janke, C.; Magiera, M.M.; Holzbaur, E.L.F. Kinesin-3 Responds to Local Microtubule Dynamics to Target Synaptic Cargo Delivery to the Presynapse. Curr. Biol. 2019, 29, 268–282.e8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  275. Grigoriev, I.; Gouveia, S.M.; van der Vaart, B.; Demmers, J.; Smyth, J.T.; Honnappa, S.; Splinter, D.; Steinmetz, M.O.; Putney, J.W.; Hoogenraad, C.C.; et al. STIM1 Is a MT-Plus-End-Tracking Protein Involved in Remodeling of the ER. Curr. Biol. 2008, 18, 177–182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  276. Honnappa, S.; Gouveia, S.M.; Weisbrich, A.; Damberger, F.F.; Bhavesh, N.S.; Jawhari, H.; Grigoriev, I.; van Rijssel, F.J.A.; Buey, R.M.; Lawera, A.; et al. An EB1-Binding Motif Acts as a Microtubule Tip Localization Signal. Cell 2009, 138, 366–376. [Google Scholar] [CrossRef] [Green Version]
  277. Asanov, A.; Sherry, R.; Sampieri, A.; Vaca, L. A Relay Mechanism between EB1 and APC Facilitate STIM1 Puncta Assembly at Endoplasmic Reticulum-Plasma Membrane Junctions. Cell Calcium 2013, 54, 246–256. [Google Scholar] [CrossRef]
  278. Decker, J.M.; Krüger, L.; Sydow, A.; Dennissen, F.J.; Siskova, Z.; Mandelkow, E.; Mandelkow, E. The Tau/A152T Mutation, a Risk Factor for Frontotemporal-spectrum Disorders, Leads to NR 2B Receptor-mediated Excitotoxicity. EMBO Rep. 2016, 17, 552–569. [Google Scholar] [CrossRef] [Green Version]
  279. Hunsberger, H.C.; Rudy, C.C.; Batten, S.R.; Gerhardt, G.A.; Reed, M.N. P301L Tau Expression Affects Glutamate Release and Clearance in the Hippocampal Trisynaptic Pathway. J. Neurochem. 2015, 132, 169–182. [Google Scholar] [CrossRef] [Green Version]
  280. Roberson, E.D.; Halabisky, B.; Yoo, J.W.; Yao, J.; Chin, J.; Yan, F.; Wu, T.; Hamto, P.; Devidze, N.; Yu, G.Q.; et al. Amyloid-β/Fyn-Induced Synaptic, Network, and Cognitive Impairments Depend on Tau Levels in Multiple Mouse Models of Alzheimer’s Disease. J. Neurosci. 2011, 31, 700–711. [Google Scholar] [CrossRef] [Green Version]
  281. Maeda, S.; Djukic, B.; Taneja, P.; Yu, G.; Lo, I.; Davis, A.; Craft, R.; Guo, W.; Wang, X.; Kim, D.; et al. Expression of A152T Human Tau Causes Age-dependent Neuronal Dysfunction and Loss in Transgenic Mice. EMBO Rep. 2016, 17, 530–551. [Google Scholar] [CrossRef] [PubMed]
  282. Sydow, A.; van der Jeugd, A.; Zheng, F.; Ahmed, T.; Balschun, D.; Petrova, O.; Drexler, D.; Zhou, L.; Rune, G.; Mandelkow, E.; et al. Tau-Induced Defects in Synaptic Plasticity, Learning, and Memory Are Reversible in Transgenic Mice after Switching Off the Toxic Tau Mutant. J. Neurosci. 2011, 31, 2511–2525. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  283. Yoshiyama, Y.; Higuchi, M.; Zhang, B.; Huang, S.-M.; Iwata, N.; Saido, T.C.; Maeda, J.; Suhara, T.; Trojanowski, J.Q.; Lee, V.M.-Y. Synapse Loss and Microglial Activation Precede Tangles in a P301S Tauopathy Mouse Model. Neuron 2007, 53, 337–351. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  284. Lopes, A.T.; Hausrat, T.J.; Heisler, F.F.; Gromova, K.V.; Lombino, F.L.; Fischer, T.; Ruschkies, L.; Breiden, P.; Thies, E.; Hermans-Borgmeyer, I.; et al. Spastin Depletion Increases Tubulin Polyglutamylation and Impairs Kinesin-Mediated Neuronal Transport, Leading to Working and Associative Memory Deficits. PLoS Biol. 2020, 18, e3000820. [Google Scholar] [CrossRef] [PubMed]
  285. Xie, J.-D.; Chen, S.-R.; Chen, H.; Zeng, W.-A.; Pan, H.-L. Presynaptic N-Methyl-d-Aspartate (NMDA) Receptor Activity Is Increased Through Protein Kinase C in Paclitaxel-Induced Neuropathic Pain. J. Biol. Chem. 2016, 291, 19364–19373. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  286. Graffe, M.; Zenisek, D.; Taraska, J.W. A Marginal Band of Microtubules Transports and Organizes Mitochondria in Retinal Bipolar Synaptic Terminals. J. Gen. Physiol. 2015, 146, 109–117. [Google Scholar] [CrossRef]
  287. Yuen, E.Y.; Jiang, Q.; Chen, P.; Gu, Z.; Feng, J.; Yan, Z. Serotonin 5-HT1A Receptors Regulate NMDA Receptor Channels through a Microtubule-Dependent Mechanism. J. Neurosci. 2005, 25, 5488–5501. [Google Scholar] [CrossRef] [Green Version]
  288. Setou, M.; Seog, D.-H.; Tanaka, Y.; Kanai, Y.; Takei, Y.; Kawagishi, M.; Hirokawa, N. Glutamate-Receptor-Interacting Protein GRIP1 Directly Steers Kinesin to Dendrites. Nature 2002, 417, 83–87. [Google Scholar] [CrossRef]
  289. Setou, M.; Nakagawa, T.; Seog, D.-H.; Hirokawa, N. Kinesin Superfamily Motor Protein KIF17 and MLin-10 in NMDA Receptor-Containing Vesicle Transport. Science 2000, 288, 1796–1802. [Google Scholar] [CrossRef]
  290. Salgado-Puga, K.; Pena-Ortega, F. Cellular and Network Mechanisms Underlying Memory Impairment Induced by Amyloid β Protein. Protein Pept. Lett. 2015, 22, 303–321. [Google Scholar] [CrossRef]
  291. Wagner, W.; Brenowitz, S.D.; Hammer, J.A. Myosin-Va Transports the Endoplasmic Reticulum into the Dendritic Spines of Purkinje Neurons. Nat. Cell Biol. 2011, 13, 40–48. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  292. McVicker, D.P.; Awe, A.M.; Richters, K.E.; Wilson, R.L.; Cowdrey, D.A.; Hu, X.; Chapman, E.R.; Dent, E.W. Transport of a Kinesin-Cargo Pair along Microtubules into Dendritic Spines Undergoing Synaptic Plasticity. Nat. Commun. 2016, 7, 12741. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  293. Shumyatsky, G.P.; Malleret, G.; Shin, R.-M.; Takizawa, S.; Tully, K.; Tsvetkov, E.; Zakharenko, S.S.; Joseph, J.; Vronskaya, S.; Yin, D.; et al. Stathmin, a Gene Enriched in the Amygdala, Controls Both Learned and Innate Fear. Cell 2005, 123, 697–709. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  294. Hou, Y.-Y.; Lu, B.; Li, M.; Liu, Y.; Chen, J.; Chi, Z.-Q.; Liu, J.-G. Involvement of Actin Rearrangements within the Amygdala and the Dorsal Hippocampus in Aversive Memories of Drug Withdrawal in Acute Morphine-Dependent Rats. J. Neurosci. 2009, 29, 12244–12254. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  295. Ahmed, T.; van der Jeugd, A.; Blum, D.; Galas, M.-C.; D’Hooge, R.; Buee, L.; Balschun, D. Cognition and Hippocampal Synaptic Plasticity in Mice with a Homozygous Tau Deletion. Neurobiol. Aging 2014, 35, 2474–2478. [Google Scholar] [CrossRef]
  296. Shipton, O.A.; Leitz, J.R.; Dworzak, J.; Acton, C.E.J.; Tunbridge, E.M.; Denk, F.; Dawson, H.N.; Vitek, M.P.; Wade-Martins, R.; Paulsen, O.; et al. Tau Protein Is Required for Amyloid-Induced Impairment of Hippocampal Long-Term Potentiation. J. Neurosci. 2011, 31, 1688–1692. [Google Scholar] [CrossRef] [Green Version]
  297. Fá, M.; Puzzo, D.; Piacentini, R.; Staniszewski, A.; Zhang, H.; Baltrons, M.A.; Li Puma, D.D.; Chatterjee, I.; Li, J.; Saeed, F.; et al. Extracellular Tau Oligomers Produce An Immediate Impairment of LTP and Memory. Sci. Rep. 2016, 6, 19393. [Google Scholar] [CrossRef] [Green Version]
  298. Hill, E.; Karikari, T.K.; Moffat, K.G.; Richardson, M.J.E.; Wall, M.J. Introduction of Tau Oligomers into Cortical Neurons Alters Action Potential Dynamics and Disrupts Synaptic Transmission and Plasticity. eNeuro 2019, 6. [Google Scholar] [CrossRef]
  299. Boekhoorn, K.; Terwel, D.; Biemans, B.; Borghgraef, P.; Wiegert, O.; Ramakers, G.J.A.; de Vos, K.; Krugers, H.; Tomiyama, T.; Mori, H.; et al. Improved Long-Term Potentiation and Memory in Young Tau-P301L Transgenic Mice before Onset of Hyperphosphorylation and Tauopathy. J. Neurosci. 2006, 26, 3514–3523. [Google Scholar] [CrossRef] [Green Version]
  300. Ondrejcak, T.; Hu, N.W.; Qi, Y.; Klyubin, I.; Corbett, G.T.; Fraser, G.; Perkinton, M.S.; Walsh, D.M.; Billinton, A.; Rowan, M.J. Soluble Tau Aggregates Inhibit Synaptic Long-Term Depression and Amyloid β-Facilitated LTD in Vivo. Neurobiol. Dis. 2019, 127, 582–590. [Google Scholar] [CrossRef]
  301. Barnes, S.J.; Opitz, T.; Merkens, M.; Kelly, T.; von der Brelie, C.; Krueppel, R.; Beck, H. Stable Mossy Fiber Long-Term Potentiation Requires Calcium Influx at the Granule Cell Soma, Protein Synthesis, and Microtubule-Dependent Axonal Transport. J. Neurosci. 2010, 30, 12996–13004. [Google Scholar] [CrossRef]
  302. Vickers, C.A.; Wyllie, D.J.A. Late-Phase, Protein Synthesis-Dependent Long-Term Potentiation in Hippocampal CA1 Pyramidal Neurones with Destabilized Microtubule Networks. Br. J. Pharmacol. 2007, 151, 1071–1077. [Google Scholar] [CrossRef] [Green Version]
  303. Regan, P.; Piers, T.; Yi, J.H.; Kim, D.H.; Huh, S.; Park, S.J.; Ryu, J.H.; Whitcomb, D.J.; Cho, K. Tau Phosphorylation at Serine 396 Residue Is Required for Hippocampal LTD. J. Neurosci. 2015, 35, 4804–4812. [Google Scholar] [CrossRef] [Green Version]
  304. Kimura, T.; Whitcomb, D.J.; Jo, J.; Regan, P.; Piers, T.; Heo, S.; Brown, C.; Hashikawa, T.; Murayama, M.; Seok, H.; et al. Microtubule-Associated Protein Tau Is Essential for Long-Term Depression in the Hippocampus. Philos. Trans. R. Soc. B Biol. Sci. 2014, 369, 20130144. [Google Scholar] [CrossRef] [Green Version]
  305. Scheuss, V.; Bonhoeffer, T. Function of Dendritic Spines on Hippocampal Inhibitory Neurons. Cereb. Cortex 2014, 24, 3142–3153. [Google Scholar] [CrossRef] [Green Version]
  306. Brandt, R.; Paululat, A. Microcompartments in the Drosophila Heart and the Mammalian Brain: General Features and Common Principles. Biol. Chem. 2013, 394, 217–230. [Google Scholar] [CrossRef]
  307. Matus, A.; Ackermann, M.; Pehling, G.; Byers, H.R.; Fujiwara, K. High Actin Concentrations in Brain Dendritic Spines and Postsynaptic Densities. Proc. Natl. Acad. Sci. USA 1982, 79, 7590–7594. [Google Scholar] [CrossRef] [Green Version]
  308. Gray, E.G.; Westrum, L.E.; Burgoyne, R.D.; Barron, J. Synaptic Organisation and Neuron Microtubule Distribution. Cell Tissue Res. 1982, 226, 579–588. [Google Scholar] [CrossRef]
  309. Westrum, L.E.; Jones, D.H.; Gray, E.G.; Barron, J. Microtubules, Dendritic Spines and Spine Apparatuses. Cell Tissue Res. 1980, 208, 171–181. [Google Scholar] [CrossRef]
  310. Conde, C.; Cáceres, A. Microtubule Assembly, Organization and Dynamics in Axons and Dendrites. Nat. Rev. Neurosci. 2009, 10, 319–332. [Google Scholar] [CrossRef]
  311. Hoogenraad, C.C.; Bradke, F. Control of Neuronal Polarity and Plasticity—A Renaissance for Microtubules? Trends Cell Biol. 2009, 19, 669–676. [Google Scholar] [CrossRef]
  312. Westrum, L.E.; Gray, E.G.; Burgoyne, R.D.; Barron, J. Synaptic Development and Microtubule Organization. Cell Tissue Res. 1983, 231, 93–102. [Google Scholar] [CrossRef]
  313. Müller-Thomsen, L.; Borgmann, D.; Morcinek, K.; Schröder, S.; Dengler, B.; Moser, N.; Neumaier, F.; Schneider, T.; Schröder, H.; Huggenberger, S. Consequences of Hyperphosphorylated Tau on the Morphology and Excitability of Hippocampal Neurons in Aged Tau Transgenic Mice. Neurobiol. Aging 2020, 93, 109–123. [Google Scholar] [CrossRef]
  314. Gu, J.; Firestein, B.L.; Zheng, J.Q. Microtubules in Dendritic Spine Development. J. Neurosci. 2008, 28, 12120–12124. [Google Scholar] [CrossRef]
  315. Geraldo, S.; Khanzada, U.K.; Parsons, M.; Chilton, J.K.; Gordon-Weeks, P.R. Targeting of the F-Actin-Binding Protein Drebrin by the Microtubule plus-Tip Protein EB3 Is Required for Neuritogenesis. Nat. Cell Biol. 2008, 10, 1181–1189. [Google Scholar] [CrossRef]
  316. Gordon-Weeks, P.R. The Role of the Drebrin/EB3/Cdk5 Pathway in Dendritic Spine Plasticity, Implications for Alzheimer’s Disease. Brain Res. Bull. 2016, 126, 293–299. [Google Scholar] [CrossRef] [Green Version]
  317. Pchitskaya, E.; Kraskovskaya, N.; Chernyuk, D.; Popugaeva, E.; Zhang, H.; Vlasova, O.; Bezprozvanny, I. Stim2-Eb3 Association and Morphology of Dendritic Spines in Hippocampal Neurons. Sci. Rep. 2017, 7, 17625. [Google Scholar] [CrossRef] [Green Version]
  318. Matsuzaki, M.; Honkura, N.; Ellis-Davies, G.C.R.; Kasai, H. Structural Basis of Long-Term Potentiation in Single Dendritic Spines. Nature 2004, 429, 761–766. [Google Scholar] [CrossRef]
  319. Okamoto, K.-I.; Nagai, T.; Miyawaki, A.; Hayashi, Y. Rapid and Persistent Modulation of Actin Dynamics Regulates Postsynaptic Reorganization Underlying Bidirectional Plasticity. Nat. Neurosci. 2004, 7, 1104–1112. [Google Scholar] [CrossRef]
  320. Hu, X.; Ballo, L.; Pietila, L.; Viesselmann, C.; Ballweg, J.; Lumbard, D.; Stevenson, M.; Merriam, E.; Dent, E.W. BDNF-Induced Increase of PSD-95 in Dendritic Spines Requires Dynamic Microtubule Invasions. J. Neurosci. 2011, 31, 15597–15603. [Google Scholar] [CrossRef] [Green Version]
  321. Caceres, A.; Payne, M.R.; Binder, L.I.; Steward, O. Immunocytochemical Localization of Actin and Microtubule-Associated Protein MAP2 in Dendritic Spines. Proc. Natl. Acad. Sci. USA 1983, 80, 1738–1742. [Google Scholar] [CrossRef] [Green Version]
  322. Buddle, M.; Eberhardt, E.; Ciminello, L.H.; Levin, T.; Wing, R.; DiPasquale, K.; Raley-Susman, K.M. Microtubule-Associated Protein 2 Associates with the NMDA Receptor and Is Spatially Redistributed within Rat Hippocampal Neurons after Oxygen-Glucose Deprivation. Brain Res. 2003, 978, 38–50. [Google Scholar] [CrossRef]
  323. Gertz, H.J.; Cervos-Navarro, J.; Ewald, V. The Septo-Hippocampal Pathway in Patients Suffering from Senile Dementia of Alzheimer’s Type: Evidence for Neuronal Plasticity? Neurosci. Lett. 1987, 76, 228–232. [Google Scholar] [CrossRef]
  324. Androuin, A.; Potier, B.; Nägerl, U.V.; Cattaert, D.; Danglot, L.; Thierry, M.; Youssef, I.; Triller, A.; Duyckaerts, C.; el Hachimi, K.H.; et al. Evidence for Altered Dendritic Spine Compartmentalization in Alzheimer’s Disease and Functional Effects in a Mouse Model. Acta Neuropathol. 2018, 135, 839–854. [Google Scholar] [CrossRef]
  325. Esteves da Silva, M.; Adrian, M.; Schätzle, P.; Lipka, J.; Watanabe, T.; Cho, S.; Futai, K.; Wierenga, C.J.; Kapitein, L.C.; Hoogenraad, C.C. Positioning of AMPA Receptor-Containing Endosomes Regulates Synapse Architecture. Cell Rep. 2015, 13, 933–943. [Google Scholar] [CrossRef] [Green Version]
  326. Glantz, L.A.; Lewis, D.A. Decreased Dendritic Spine Density on Prefrontal Cortical Pyramidal Neurons in Schizophrenia. Arch. Gen. Psychiatry 2000, 57, 65–73. [Google Scholar] [CrossRef] [Green Version]
  327. Irwin, S.A.; Patel, B.; Idupulapati, M.; Harris, J.B.; Crisostomo, R.A.; Larsen, B.P.; Kooy, F.; Willems, P.J.; Cras, P.; Kozlowski, P.B.; et al. Abnormal Dendritic Spine Characteristics in the Temporal and Visual Cortices of Patients with Fragile-X Syndrome: A Quantitative Examination. Am. J. Med. Genet. 2001, 98, 161–167. [Google Scholar] [CrossRef]
  328. Penazzi, L.; Tackenberg, C.; Ghori, A.; Golovyashkina, N.; Niewidok, B.; Selle, K.; Ballatore, C.; Smith, A.B.; Bakota, L.; Brandt, R. Aβ-Mediated Spine Changes in the Hippocampus Are Microtubule-Dependent and Can Be Reversed by a Subnanomolar Concentration of the Microtubule-Stabilizing Agent Epothilone D. Neuropharmacology 2016, 105, 84–95. [Google Scholar] [CrossRef] [Green Version]
  329. Chuckowree, J.A.; Vickers, J.C. Cytoskeletal and Morphological Alterations Underlying Axonal Sprouting after Localized Transection of Cortical Neuron Axons In Vitro. J. Neurosci. 2003, 23, 3715–3725. [Google Scholar] [CrossRef] [Green Version]
  330. Tang-Schomer, M.D.; Johnson, V.E.; Baas, P.W.; Stewart, W.; Smith, D.H. Partial Interruption of Axonal Transport Due to Microtubule Breakage Accounts for the Formation of Periodic Varicosities after Traumatic Axonal Injury. Exp. Neurol. 2012, 233, 364–372. [Google Scholar] [CrossRef] [Green Version]
  331. Qu, X.; Yuan, F.N.; Corona, C.; Pasini, S.; Pero, M.E.; Gundersen, G.G.; Shelanski, M.L.; Bartolini, F. Stabilization of Dynamic Microtubules by MDia1 Drives Tau-Dependent Aβ1–42 Synaptotoxicity. J. Cell Biol. 2017, 216, 3161–3178. [Google Scholar] [CrossRef] [Green Version]
  332. Gu, J.; Zheng, J.Q. Microtubules in Dendritic Spine Development and Plasticity. Open Neurosci. J. 2009, 3, 128–133. [Google Scholar] [CrossRef]
  333. Briones, T.L.; Woods, J. Chemotherapy-Induced Cognitive Impairment Is Associated with Decreases in Cell Proliferation and Histone Modifications. BMC Neurosci. 2011, 12, 124. [Google Scholar] [CrossRef] [Green Version]
  334. Winocur, G.; Wojtowicz, J.M.; Tannock, I.F. Memory Loss in Chemotherapy-Treated Rats Is Exacerbated in High-Interference Conditions and Related to Suppression of Hippocampal Neurogenesis. Behav. Brain Res. 2015, 281, 239–244. [Google Scholar] [CrossRef]
  335. Muallaoglu, S.; Disel, U.; Mertsoylu, H.; Besen, A.; Karadeniz, C.; Sumbul, A.T.; Abali, H.; Ozyilkan, O. Acute Infusion Reactions to Chemotherapeutic Drugs: A Single Institute Experience. J. BUON 2013, 18, 261–267. [Google Scholar]
  336. Perry, J.R.; Warner, E. Transient Encephalopathy after Paclitaxel (Taxol) Infusion. Neurology 1996, 46, 1596–1599. [Google Scholar] [CrossRef]
  337. Yousefzadeh, S.A.; Youngkin, A.E.; Lusk, N.A.; Wen, S.; Meck, W.H. Bidirectional Role of Microtubule Dynamics in the Acquisition and Maintenance of Temporal Information in Dorsolateral Striatum. Neurobiol. Learn. Mem. 2021, 183, 107468. [Google Scholar] [CrossRef]
  338. Cassar, M.; Law, A.D.; Chow, E.S.; Giebultowicz, J.M.; Kretzschmar, D. Disease-Associated Mutant Tau Prevents Circadian Changes in the Cytoskeleton of Central Pacemaker Neurons. Front. Neurosci. 2020, 14, 232. [Google Scholar] [CrossRef]
  339. Mershin, A.; Pavlopoulos, E.; Fitch, O.; Braden, B.C.; Nanopoulos, D.V.; Skoulakis, E.M.C. Learning and Memory Deficits Upon TAU Accumulation in Drosophila Mushroom Body Neurons. Learn. Mem. 2004, 11, 277–287. [Google Scholar] [CrossRef] [Green Version]
  340. Pennanen, L.; Wolfer, D.P.; Nitsch, R.M.; Gotz, J. Impaired Spatial Reference Memory and Increased Exploratory Behavior in P301L Tau Transgenic Mice. Genes Brain Behav. 2006, 5, 369–379. [Google Scholar] [CrossRef]
  341. Uchida, S.; Teubner, B.J.W.; Hevi, C.; Hara, K.; Kobayashi, A.; Dave, R.M.; Shintaku, T.; Jaikhan, P.; Yamagata, H.; Suzuki, T.; et al. CRTC1 Nuclear Translocation Following Learning Modulates Memory Strength via Exchange of Chromatin Remodeling Complexes on the Fgf1 Gene. Cell Rep. 2017, 18, 352–366. [Google Scholar] [CrossRef]
  342. Muhia, M.; Thies, E.; Labonté, D.; Ghiretti, A.E.; Gromova, K.V.; Xompero, F.; Lappe-Siefke, C.; Hermans-Borgmeyer, I.; Kuhl, D.; Schweizer, M.; et al. The Kinesin KIF21B Regulates Microtubule Dynamics and Is Essential for Neuronal Morphology, Synapse Function, and Learning and Memory. Cell Rep. 2016, 15, 968–977. [Google Scholar] [CrossRef] [Green Version]
  343. Ballatore, C.; Brunden, K.R.; Huryn, D.M.; Trojanowski, J.Q.; Lee, V.M.Y.; Smith, A.B. Microtubule Stabilizing Agents as Potential Treatment for Alzheimers Disease and Related Neurodegenerative Tauopathies. J. Med. Chem. 2012, 55, 8979–8996. [Google Scholar] [CrossRef] [Green Version]
  344. Guo, B.; Huang, Y.; Gao, Q.; Zhou, Q. Stabilization of Microtubules Improves Cognitive Functions and Axonal Transport of Mitochondria in Alzheimer’s Disease Model Mice. Neurobiol. Aging 2020, 96, 223–232. [Google Scholar] [CrossRef]
  345. Cross, D.J.; Meabon, J.S.; Cline, M.M.; Richards, T.L.; Stump, A.J.; Cross, C.G.; Minoshima, S.; Banks, W.A.; Cook, D.G. Paclitaxel Reduces Brain Injury from Repeated Head Trauma in Mice. J. Alzheimers Dis. 2019, 67, 859–874. [Google Scholar] [CrossRef]
  346. Chang, A.; Chung, N.-C.; Lawther, A.J.; Ziegler, A.I.; Shackleford, D.M.; Sloan, E.K.; Walker, A.K. The Anti-Inflammatory Drug Aspirin Does Not Protect Against Chemotherapy-Induced Memory Impairment by Paclitaxel in Mice. Front. Oncol. 2020, 10, 564965. [Google Scholar] [CrossRef]
  347. Fardell, J.E.; Vardy, J.; Johnston, I.N. The Short and Long Term Effects of Docetaxel Chemotherapy on Rodent Object Recognition and Spatial Reference Memory. Life Sci. 2013, 93, 596–604. [Google Scholar] [CrossRef]
  348. Callaghan, C.K.; O’Mara, S.M. Long-Term Cognitive Dysfunction in the Rat Following Docetaxel Treatment Is Ameliorated by the Phosphodiesterase-4 Inhibitor, Rolipram. Behav. Brain Res. 2015, 290, 84–89. [Google Scholar] [CrossRef]
  349. Atarod, D.; Eskandari-Sedighi, G.; Pazhoohi, F.; Karimian, S.M.; Khajeloo, M.; Riazi, G.H. Microtubule Dynamicity Is More Important than Stability in Memory Formation: An In Vivo Study. J. Mol. Neurosci. 2015, 56, 313–319. [Google Scholar] [CrossRef]
  350. Panoz-Brown, D.; Carey, L.M.; Smith, A.E.; Gentry, M.; Sluka, C.M.; Corbin, H.E.; Wu, J.-E.; Hohmann, A.G.; Crystal, J.D. The Chemotherapeutic Agent Paclitaxel Selectively Impairs Reversal Learning While Sparing Prior Learning, New Learning and Episodic Memory. Neurobiol. Learn. Mem. 2017, 144, 259–270. [Google Scholar] [CrossRef] [Green Version]
  351. Huehnchen, P.; Boehmerle, W.; Springer, A.; Freyer, D.; Endres, M. A Novel Preventive Therapy for Paclitaxel-Induced Cognitive Deficits: Preclinical Evidence from C57BL/6 Mice. Transl. Psychiatry 2017, 7, e1185. [Google Scholar] [CrossRef] [Green Version]
  352. Fardell, J.E.; Zhang, J.; de Souza, R.; Vardy, J.; Johnston, I.; Allen, C.; Henderson, J.; Piquette-Miller, M. The Impact of Sustained and Intermittent Docetaxel Chemotherapy Regimens on Cognition and Neural Morphology in Healthy Mice. Psychopharmacology 2014, 231, 841–852. [Google Scholar] [CrossRef]
  353. Li, Z.; Zhao, S.; Zhang, H.-L.; Liu, P.; Liu, F.-F.; Guo, Y.-X.; Wang, X.-L. Proinflammatory Factors Mediate Paclitaxel-Induced Impairment of Learning and Memory. Mediat. Inflamm. 2018, 2018, 3941840. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  354. Nguyen, L.D.; Fischer, T.T.; Ehrlich, B.E. Pharmacological Rescue of Cognitive Function in a Mouse Model of Chemobrain. Mol. Neurodegener. 2021, 16, 41. [Google Scholar] [CrossRef]
  355. Seigers, R.; Loos, M.; van Tellingen, O.; Boogerd, W.; Smit, A.B.; Schagen, S.B. Cognitive Impact of Cytotoxic Agents in Mice. Psychopharmacology 2015, 232, 17–37. [Google Scholar] [CrossRef]
  356. Bensimon, G.; Chermat, R. Microtubule Disruption and Cognitive Defects: Effect of Colchicine on Learning Behavior in Rats. Pharmacol. Biochem. Behav. 1991, 38, 141–145. [Google Scholar] [CrossRef]
  357. Nakayama, T.; Sawada, T. Involvement of Microtubule Integrity in Memory Impairment Caused by Colchicine. Pharmacol. Biochem. Behav. 2002, 71, 119–138. [Google Scholar] [CrossRef]
  358. Di Patre, P.L.; Oh, J.D.; Simmons, J.M.; Butcher, L.L. Intrafimbrial Colchicine Produces Transient Impairment of Radial-Arm Maze Performance Correlated with Morphologic Abnormalities of Septohippocampal Neurons Expressing Cholinergic Markers and Nerve Growth Factor Receptor. Brain Res. 1990, 523, 316–320. [Google Scholar] [CrossRef]
  359. Mileusnic, R.; Lancashire, C.L.; Rose, S.P.R. Recalling an Aversive Experience by Day-Old Chicks Is Not Dependent on Somatic Protein Synthesis. Learn. Mem. 2005, 12, 615–619. [Google Scholar] [CrossRef] [Green Version]
  360. Kumar, A.; Seghal, N.; S Naidu, P.S.; Padi, S.S.; Goyal, R. Colchicines-Induced Neurotoxicity as an Animal Model of Sporadic Dementia of Alzheimer’s Type. Pharmacol. Rep. 2007, 59, 274–283. [Google Scholar]
  361. Tilson, H.A.; Rogers, B.C.; Grimes, L.; Harry, G.J.; Peterson, N.J.; Hong, J.S.; Dyer, R.S. Time-Dependent Neurobiological Effects of Colchicine Administered Directly into the Hippocampus of Rats. Brain Res. 1987, 408, 163–172. [Google Scholar] [CrossRef]
  362. Tilson, H.A.; Peterson, N.J. Colchicine as an Investigative Tool in Neurobiology. Toxicology 1987, 46, 159–175. [Google Scholar] [CrossRef]
  363. Walsh, T.J.; Schulz, D.W.; Tilson, H.A.; Schmechel, D.E. Cochicine-Induced Granule Cell Loss in Rat Hippocampus: Selective Behavioral and Histological Alterations. Brain Res. 1986, 398, 23–36. [Google Scholar] [CrossRef]
  364. Goldschmidt, R.B.; Steward, O. Neurotoxic Effects of Colchicine: Differential Susceptibility of CNS Neuronal Populations. Neuroscience 1982, 7, 695–714. [Google Scholar] [CrossRef]
  365. Lothman, E.W.; Stein, D.A.; Wooten, G.F.; Zucker, D.K. Potential Mechanisms Underlying the Destruction of Dentate Gyrus Granule Cells by Colchicine. Exp. Neurol. 1982, 78, 293–302. [Google Scholar] [CrossRef]
  366. Jarrard, L.E.; Okaichi, H.; Steward, O.; Goldschmidt, R.B. On the Role of Hippocampal Connections in the Performance of Place and Cue Tasks: Comparisons with Damage to Hippocampus. Behav. Neurosci. 1984, 98, 946–954. [Google Scholar] [CrossRef]
  367. Ionescu, T. Exploring the nature of cognitive flexibility. New Ideas Psychol. 2012, 30, 190–200. [Google Scholar] [CrossRef]
  368. Tanimura, Y.; Yang, M.C.; Lewis, M.H. Procedural learning and cognitive flexibility in a mouse model of restricted, repetitive behaviour. Behav. Brain Res. 2008, 189, 250–256. [Google Scholar] [CrossRef]
  369. Cowen, S.L.; Phelps, C.E.; Navratilova, E.; McKinzie, D.L.; Okun, A.; Husain, O.; Gleason, S.D.; Witkin, J.M.; Porreca, F. Chronic pain impairs cognitive flexibility and engages novel learning strategies in rats. Pain 2018, 159, 1403–1412. [Google Scholar] [CrossRef]
Figure 1. Microtubules (MTs) function in different neuronal compartments. In physiological conditions (A), MTs allow synaptic vesicles’ transport and recycling for proper synaptic transmission and plasticity. MTs contribute to the organization of the axon initial segment for action potential initiation and plasticity of intrinsic excitability. MTs are also capable of invading dendritic spines in an activity-dependent manner, for cargo delivery and postsynaptic plasticity regulation. However, when MTs dynamics and stability processes exceed the normal homeostatic range, under pathological conditions (B), there are MTs-dependent alterations in synaptic transmission due to inefficient synaptic vesicles’ transport, aberrant firing activity due to the relocation of the axon initial segment, loss of postsynaptic plasticity and alterations in dendritic spines due to reduced responsivity of MTs located in the dendritic shaft.
Figure 1. Microtubules (MTs) function in different neuronal compartments. In physiological conditions (A), MTs allow synaptic vesicles’ transport and recycling for proper synaptic transmission and plasticity. MTs contribute to the organization of the axon initial segment for action potential initiation and plasticity of intrinsic excitability. MTs are also capable of invading dendritic spines in an activity-dependent manner, for cargo delivery and postsynaptic plasticity regulation. However, when MTs dynamics and stability processes exceed the normal homeostatic range, under pathological conditions (B), there are MTs-dependent alterations in synaptic transmission due to inefficient synaptic vesicles’ transport, aberrant firing activity due to the relocation of the axon initial segment, loss of postsynaptic plasticity and alterations in dendritic spines due to reduced responsivity of MTs located in the dendritic shaft.
Cells 11 00923 g001
Figure 2. Differential effects caused by Paclitaxel. Paclitaxel is a microtubules stabilizer used in chemotherapy, which is capable of exerting both negative and positive effects on neural function and shape. (A) The most common negative effects associated with PTX are neuronal death, cognitive impairment, and sensory abnormalities (i.e., allodynia), that correlate with changes in brain activity and connectivity. (B) On the other hand, some positive effects of PTX are the enhancement of regeneration and elongation of axonal processes, restoration of axonal transport, and the reduction of glutamate-induced neurotoxicity.
Figure 2. Differential effects caused by Paclitaxel. Paclitaxel is a microtubules stabilizer used in chemotherapy, which is capable of exerting both negative and positive effects on neural function and shape. (A) The most common negative effects associated with PTX are neuronal death, cognitive impairment, and sensory abnormalities (i.e., allodynia), that correlate with changes in brain activity and connectivity. (B) On the other hand, some positive effects of PTX are the enhancement of regeneration and elongation of axonal processes, restoration of axonal transport, and the reduction of glutamate-induced neurotoxicity.
Cells 11 00923 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Peña-Ortega, F.; Robles-Gómez, Á.A.; Xolalpa-Cueva, L. Microtubules as Regulators of Neural Network Shape and Function: Focus on Excitability, Plasticity and Memory. Cells 2022, 11, 923. https://doi.org/10.3390/cells11060923

AMA Style

Peña-Ortega F, Robles-Gómez ÁA, Xolalpa-Cueva L. Microtubules as Regulators of Neural Network Shape and Function: Focus on Excitability, Plasticity and Memory. Cells. 2022; 11(6):923. https://doi.org/10.3390/cells11060923

Chicago/Turabian Style

Peña-Ortega, Fernando, Ángel Abdiel Robles-Gómez, and Lorena Xolalpa-Cueva. 2022. "Microtubules as Regulators of Neural Network Shape and Function: Focus on Excitability, Plasticity and Memory" Cells 11, no. 6: 923. https://doi.org/10.3390/cells11060923

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