Next Article in Journal
A Brief Review on the Biology and Effects of Cellular and Circulating microRNAs on Cardiac Remodeling after Infarction
Next Article in Special Issue
The Association between α-Synuclein and α-Tubulin in Brain Synapses
Previous Article in Journal
Solvates of New Arylpiperazine Salicylamide Derivative-a Multi-Technique Approach to the Description of 5 HTR Ligand Structure and Interactions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Neurons and Glia Interplay in α-Synucleinopathies

Center of Clinical Research, Experimental Surgery and Translational Research, Biomedical Research Foundation of the Academy of Athens, 11527 Athens, Greece
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(9), 4994; https://doi.org/10.3390/ijms22094994
Submission received: 6 April 2021 / Revised: 29 April 2021 / Accepted: 4 May 2021 / Published: 8 May 2021
(This article belongs to the Special Issue α-Synuclein in Neurons and Glia: From Physiology to Pathology)

Abstract

:
Accumulation of the neuronal presynaptic protein alpha-synuclein within proteinaceous inclusions represents the key histophathological hallmark of a spectrum of neurodegenerative disorders, referred to by the umbrella term a-synucleinopathies. Even though alpha-synuclein is expressed predominantly in neurons, pathological aggregates of the protein are also found in the glial cells of the brain. In Parkinson’s disease and dementia with Lewy bodies, alpha-synuclein accumulates mainly in neurons forming the Lewy bodies and Lewy neurites, whereas in multiple system atrophy, the protein aggregates mostly in the glial cytoplasmic inclusions within oligodendrocytes. In addition, astrogliosis and microgliosis are found in the synucleinopathy brains, whereas both astrocytes and microglia internalize alpha-synuclein and contribute to the spread of pathology. The mechanisms underlying the pathological accumulation of alpha-synuclein in glial cells that under physiological conditions express low to non-detectable levels of the protein are an area of intense research. Undoubtedly, the presence of aggregated alpha-synuclein can disrupt glial function in general and can contribute to neurodegeneration through numerous pathways. Herein, we summarize the current knowledge on the role of alpha-synuclein in both neurons and glia, highlighting the contribution of the neuron-glia connectome in the disease initiation and progression, which may represent potential therapeutic target for a-synucleinopathies.

1. Introduction

The presynaptic neuronal protein alpha-synuclein (aSyn) under physiological conditions regulates neurotransmitter release and SNARE (soluble N-ethylmaleimide-sensitive factor attachment protein receptor) complex assembly and is considered a chameleon-protein due to its remarkable conformational plasticity [1]. On the other hand, aggregated aSyn is the major component of the proteinaceous inclusions found in the degenerating neurons of Parkinson’s disease (PD) and dementia with Lewy bodies (DLB) brains, known as Lewy bodies (LBs) and Lewy neurites (LNs) [2]. Alpha-synuclein also has a strong genetic link to PD pathogenesis, since missense point mutations of the SNCA gene encoding for aSyn, SNCA gene locus duplications and triplications or gene-enhanced expression are the main causes of familial PD [3,4,5,6,7,8,9]. In contrast, multiple system atrophy (MSA), a fatal debilitating neurodegenerative disorder, is characterized by the presence of aggregated aSyn within the glial cytoplasmic inclusions (GCIs) present in the cytoplasm of oligodendrocytes [10,11,12]. Glial aSyn accumulation is also evident in PD and PD with aSyn-positive deposits reported in astrocytes and oligodendrocytes [13,14,15]. Contrarily, aSyn-positive inclusions in astrocytes have been also found in MSA [16], but to a lesser extent [14] compared to neuronal and oligodendroglial inclusion pathology.
The clinical and neuropathological heterogeneity in a-synucleinopathies may ascend from the unique properties of the different conformational aSyn strains found in neurons or glia that might contribute to distinct clinical phenotypes [17,18,19,20]. Even though the physiological and pathological functions of aSyn in neurons, where the protein is physiologically expressed, are well characterized, the mechanisms underlying the pathological accumulation of aSyn in the glial cells of the central nervous system (CNS) still necessitates further investigation. Microglia and astroglia have distinct roles in maintaining brain’s homeostasis but under stress conditions, such as increased aSyn burden, they can become activated and contribute to disease pathology by triggering neuroinflammatory mechanisms. Reactive astrocytes and microglia have been detected in human post-mortem brains of a-synucleinopathies [13,21,22,23,24,25], further supporting a role of active gliosis in the initiation and progression of the disease. Moreover, all glial cells have been reported to internalize aSyn and a neuron-to-glia transmission is thought to underlie the propagation of aSyn pathology in a-synucleinopathies. In the following sections, we discuss how aSyn affects neuronal and glial function and homeostasis in health and disease.

2. Alpha-Synuclein in Neurons: A Multifaceted Protein

2.1. A Role at the Synapse

alpha-Synuclein (aSyn) is a small, intrinsically disordered protein that is mainly localized at the pre-synaptic terminal [26,27], but is also present in the neuronal somato-dendritic compartment [28], in red blood cells [29], in the gut and other peripheral tissues [30,31,32]. Although aSyn is highly enriched in presynaptic boutons, it displays a delayed distribution in the terminals, suggesting that it is implicated in later stages of synaptic development, rather than playing a central role in synapse modulation [27]. Importantly, aSyn is differentially expressed in the various neuronal cell types, being more abundant in excitatory synapses across different brain regions and particularly in central catecholaminergic systems [33]. On the contrary, the protein displays a differential expression profile in inhibitory synapses amongst the different brain areas, with a particular interest of aSyn presence in striatal GABAergic medium spiny neurons (MSNs) [34,35].
The first indication regarding the role of aSyn on neural plasticity arose about 25 years ago, when “synelfin” (synuclein, NACP) expression was found up-regulated during bird song learning [36]. The localization of aSyn in pre-synaptic boutons is mainly attributed to its tight association with synaptic vesicle membranes [37] and its high affinity for the SNARE complex proteins synaptobrevin-2 (or Vesicle Associated Membrane Protein 2, VAMP2), synapsin III and rab3A [38,39,40]. It has been proposed that aSyn interacts with VAMP2 and promotes SNARE complex assembly [38], followed then by its disassembly in order to complete the round of membrane fusion (Figure 1). The crucial role of aSyn assembly with SNARE complex on neuronal survival was further verified by the neuronal dysfunction and impaired survival of triple αβγ-synuclein knockout mice during ageing [38,41]. Interestingly, aSyn lentiviral overexpression in primary neurons led to enhanced SNARE complex assembly, further supporting the role of this protein in synaptic activity [38]. The same group later showed that only multimeric membrane-bound, but not the soluble monomeric aSyn, can promote the SNARE complex assembly [42]. It has been also recently suggested that aSyn is involved in synaptic vesicle homeostasis at the pre-synaptic terminal via a calcium (Ca2+)-dependent mechanism [43].
On the contrary, natively unfolded monomeric aSyn at the pre-synaptic terminal is prone to form pathological conformations, thus exerting neurotoxic effects [44] (Figure 1). It has been additionally suggested that aSyn is preferably bound to synapsin 1 and VAMP2 when the protein is present in its oligomeric form [45], highlighting the importance of the conformational state of aSyn for its proper function. There are also findings supporting the implication of aSyn in synaptic transmission, due to its association with the synaptic vesicle pool, modulating the vesicle mobility, the recycling pool homeostasis and endocytosis [46,47,48].
Alpha-synuclein can also function as a molecular chaperone via effective binding to other intracellular proteins. The first indication came with the discovery that aSyn displays structural and functional homology with other molecular chaperones, as the 14-3-3 or small heat shock proteins [49,50]. Additional studies revealed that aSyn synergistically acts with the presynaptic cysteine-string protein-alpha (CSPalpha) promoting the assembly of the SNARE complex [38,51], further validating its chaperoning properties. Biochemical and structural analysis of aSyn strengthened the current indications for its chaperone-like function via its C-terminal region (residues 61-140) [52,53,54]. However, following studies indicated that the chaperone-binding site of aSyn lies within the non-amyloidal component (NAC) region (residues 61-95), which is prone to aggregation and thus highly susceptible to form fibrils [55,56].

2.2. Association with Membranes and Lipid Trafficking

Intracellular aSyn can be found either natively unfolded in a soluble state or membrane-bound forming an alpha-helical or a beta-sheet secondary structure, depending on the solution conditions [57,58,59]. It has been proposed that there is a bidirectional link between aSyn species formation and membrane remodeling, meaning that not only aSyn structure is affected upon lipid interaction, but also that membrane integrity depends on the presence of different aSyn conformations [60,61,62]. However, there are controversial results regarding the association of aSyn with membrane lipids and its conformational state, with some studies reporting that membrane-bound aSyn gets protected from aggregation, thus leading to neurotoxicity attenuation [44,63,64], whereas others suggest that interaction of aSyn with membranes triggers its self-association and subsequent aggregation [65,66,67]. Importantly, it has been shown that the PD-related aSyn mutations reduce its interaction with membranes, thus further suggesting that aSyn binding on membranes may exert neuroprotective effects [68,69,70,71,72].
A plethora of studies argue that aSyn in its soluble state exists as a monomer [73,74,75,76], whereas others suggest that it occurs physiologically as a tetramer resisting aggregation [77,78,79]. In the presence of lipid membranes, aSyn adopts an alpha helical structure in the N-terminus region that stabilizes the formation of high-order aSyn multimers [42,73,80,81]. Interestingly, the membrane curvature seems to affect the structure of aSyn, which can adopt either an elongated or a broken alpha-helix conformation, when bound to a large diameter (∼100 nm) or a small, highly curved vesicle, respectively [82,83,84,85]. It has been also proposed that aSyn has a role in lipid metabolism, since it participates in fatty acids transportation between the cytosol and membranous compartments [86,87] and in lipid and membrane biogenesis organizing and stabilizing the lipid bilayer of membranes and vesicles [88,89]. On the other hand, disrupted aSyn expression pattern leads to lipid dysregulation, since both the absence and the overexpression of either wild-type (WT) or mutated aSyn gives rise to abnormal lipid metabolism [90,91,92,93]. Finally, several studies have demonstrated that aSyn regulates membrane homeostasis via inhibition of phospholipases activity, such as phospholipase D [94,95,96,97]; however, there are controversial results in the literature [98].

2.3. Aggregation and Post-Translational Modifications

alpha-Synuclein is composed of three distinct domains: the N-terminal lipid-binding domain, the NAC region and the C-terminal binding domain [84,99,100]. A central role in the fibril formation and subsequent aggregation of aSyn is thought to be mediated through the NAC region of the protein composed of nonpolar side-chains and assembles cross b-structures. Based on that, it has been shown that the deletion of specific residues (74-84) within the core region can abolish aSyn aggregation [101,102]. It has been also demonstrated that the endogenous neuronal aSyn and the interaction of aSyn with lipids plays a central role for aSyn recruitment and subsequent seeding of pathology, as it could behave as a core for the formation of insoluble aggregates [35,75,103,104].
Several mutations in the Snca gene have been linked to PD pathogenesis, such as the A53T, A30P, E46K, H50Q, G51D, A18P, pA29S and A53E mutations, all located in the N-terminus region [3,5,7,68,105,106,107]. Most of them are tightly linked to enhanced aSyn aggregation, pathology progression and clinical manifestations in PD. Specifically, A53T and A30P aSyn mutants are natively unfolded, similarly to WT protein. However, at higher concentrations A53T has been shown to accelerate aSyn fibrillization, a critical event in PD pathogenesis [108,109,110]. On the other hand, A30P promotes aSyn oligomerization rather than fibrillization, thus reducing aggregate formation [109,111]. The E46K mutation leads to conformational changes of aSyn due to C-terminal to N-terminal contacts in the monomeric protein, resulting in enhanced aSyn accumulation [111,112,113]. Moreover, the PD-linked H50Q point mutation increases aSyn aggregation propensity and toxicity [114], whereas the G51D mutation has the opposite effects [115]. However, although G51D mutants seem to oligomerize in a slow rate, they form more toxic fibrils, thus suggesting distinct disease mechanisms for the various aSyn mutations [116,117]. Similarly, A53E mutant seems to lead to neuronal toxicity via an aSyn aggregation-independent manner [118]. Strikingly, the G51D and A53E aSyn mutations have been proposed as potential links between PD and MSA [106,119]. However, up-to-date, no hereditable mutations in the coding region of SNCA gene have been identified in MSA cases [120]. Apart from point mutations [117,121,122], various post-translational modifications are implicated in aSyn aggregation, the most important of which are phosphorylation, sumoylation, ubiquitination, nitration, N-acetylation, O-GlcNAcylation and truncation.
The phosphorylation of aSyn both at serine and tyrosine residues and particularly at Ser129 is widely considered as an indicator of pathology. However, the effect of Ser129 phosphorylation on aSyn toxicity is still under debate, with the majority of studies suggesting that it accelerates cell toxicity and neurodegeneration [123,124,125,126,127]. Contrarily, others have proposed a neuroprotective role of Ser129 phosphorylation since it was reported to drive the conversion of toxic oligomers into less harmful aggregates [128,129,130]. Other mechanisms of phosphorylated Ser129 aSyn-mediated neuroprotection include inhibition of its fibrillation [131], upregulation of tyrosine hydroxylase (TH) activity [132] or lowering of the protein’s membrane-binding affinity [133]. Although the 90% of aSyn in LBs is found phosphorylated at Ser129, a significant amount of phosphorylated Ser129 aSyn is also detected in a soluble, rather than in an aggregated state in PD brains [134], whereas only a small percentage of aSyn is phosphorylated at Ser129 in the brains of healthy controls [135,136,137]. In addition, aSyn can be phosphorylated at Ser87, Tyr125, Tyr133 and Tyr136 residues [138,139] and these are also implicated in either neurotoxic or neuroprotective events [127,138,140,141]. Nonetheless, in most in vivo models where aSyn is overexpressed (virally, transgenic or PFF-inoculations) the detection of pSer129 positive aSyn signal is invariably linked to neurotoxicity, indicating a rather neurotoxic and not a neuroprotective role.
Nitrated aSyn is also tightly linked to neurodegeneration, as demonstrated by experiments in both cellular and animal models, as well as in patient-derived brains [142,143,144,145], through its implication in oxidative damage and disease development [146]. Four tyrosine residues in aSyn sequence, Tyr39 (within the N-terminus), Tyr125, Tyr133 and Tyr 136 (within the C-terminus) can undergo nitration. Nitration at Tyr39 has been shown to result in low binding affinity of aSyn on lipid vesicles due to its loss-of-alpha helical conformation status [147], whereas nitration at Tyr125 seems to play a crucial role for aSyn dimerization [148]. Moreover, the linking between two tyrosines is considered as a potential mechanism for aSyn oligomer stabilization and its subsequent aggregation into proteinaceous inclusions [149]. In addition, the detection of nitrated aSyn in the human blood serum could potentially serve as a clinical biomarker for PD diagnosis [150].
Another aSyn post-translational modification crucial for its aggregation propensity is ubiquitination, via regulation of the proteasome-dependent protein degradation [151] and the subcellular localization of the protein [152]. Ubiquitinated aSyn has been isolated from LBs and sarkosyl-insoluble fractions derived from synucleinopathy brains [153,154]. CHIP (C-terminal U-box domain of co-chaperone Hsp70-interacting protein), SIAH (seven in absentia homolog) and Nedd4 (neuronal precursor cell-expressed, developmentally down-regulated gene 4) have been identified among the E3 ubiquitin ligases implicated in aSyn ubiquitination [155,156,157,158,159,160]. Ubiquitin modification has been demonstrated to have differential effects on aSyn accumulation and subsequent aggregation, dependent on the residue being modified. More precisely, ubiquitination at Lys6, Lys12 and Lys21 residues has been shown to moderately inhibit aSyn fibrillation, whereas at Lys10 and Lys23 residues has been reported to promote the formation of aSyn inclusions [161]. In addition, ubiquitination at Lys32, Lys34, Lys43 and Lys96 inhibits aSyn aggregation [161].
Sumoylation is a similar process to ubiquitination, since aSyn is conjugated to SUMO (small ubiquitin-like modifier) at lysine residues. SUMO-1 was found in aSyn-positive inclusions of a-synucleinopathy brains or associated with lysosomes of PD animal models [162,163,164]. It has been also suggested that aSyn sumoylation facilitates its aggregation since it inhibits its degradation [165], whereas other studies proposed a neuroprotective role of aSyn sumoylation, which seems to promote aSyn solubility and thus inhibit its aggregation [166,167]. The discrepancy between these data may be attributed to the different lysine residues available for sumoylation being investigated in each study. Another aSyn modification that has been up for debate is its N-terminal acetylation. Although many studies have assigned a neurotoxic role on aSyn N-acetylation, as it has been shown to promote aSyn β-sheet formation and fibrillation [168,169,170], others suggest that either N-acetylated aSyn mediates its physiological binding on synaptic vesicles [171], or it acts in a protective manner against aSyn aggregation [172,173].
O-GlcNAcylation is a biochemical process that involves the attachment of O-linked N-acetylglucosamine to Ser and Thr residues of various proteins, amongst which is aSyn. Murine and human aSyn have been shown to be O-GlcNAcyled in many threonine residues including Thr33, Thr34, Thr54, Thr59, Thr64, Thr72, Thr75, Thr81 and Thr87 [174,175,176,177,178] and this post-translational modification has repetitively been linked to reduced aSyn aggregation and attenuation of PD-related toxicity [179,180,181,182]. Finally, aSyn truncation has gained scientific attention, given that C-terminally truncated aSyn has been identified in the inclusions present in PD brains [183,184,185]. Many studies have considered that aSyn truncations have neurotoxic effects due to increased accumulation of misfolded aSyn [186,187,188,189,190,191,192,193,194].

2.4. Channel Formation/Channel Interactions

As mentioned above, membrane-bound aSyn adopts an alpha-helical conformation, which facilitates its oligomerization and subsequent aggregation. It has been suggested that aSyn oligomers can form transmembrane channels and pore-like structures that have been linked to pathological events during PD development (Figure 1) [195,196,197]. As a result, vesicles or low-molecular mass molecules may penetrate the cell membrane and in combination with altered cellular ionic homeostasis could potentially lead to cell toxicity and neuronal degeneration [198,199]. Another mechanism for the increased membrane permeability involves the incorporation of aSyn oligomers between the membrane phospholipids, thus leading to the bilayer thinning which thereafter allows the diffusion of small molecules [200].
A wide range of studies has demonstrated that the ion channels formed by oligomeric aSyn dysregulate cellular ion concentrations and may represent a critical event in the pathogenesis of a-synucleinopathies [198]. Some PD-linked aSyn mutations, such as E46K and A53T, have been shown to be implicated in the channel formation, whereas other aSyn mutants (i.e., A30P) have displayed low membrane affinity [197,201]. However, other groups have shown that A30P and A53T aSyn mutations are responsible for the formation of large membrane pores through which most cations can pass non-selectively [202]. It has been reported that the formation of such cation-permeable pores could lead either to ion conductivity or to increased Ca2+ influx and subsequent cell death [202,203,204,205]. Upon aSyn cation channel opening, other channels, such as the ATP-dependent potassium channels K (ATP), have been reported to be activated in hippocampal neurons and this could probably diminish the aSyn-dependent neuronal excitability [205].
Binding of aSyn to the plasma membrane results in the formation of aggregates and this aggregation leads to the redistribution of the α3 subunit of Na+/K+-ATPase. As a result, Na+/K+-ATPase is no longer able to effectively pump out Na+ from neurons, thus leading to an intracellular Na+ accumulation [206]. Furthermore, extracellular aSyn was reported to activate the voltage-gated Ca2+ channel Cav2.2 in rat neurons, due to disorganization of lipid rafts in the plasma membrane, resulting in enhanced dopamine release and increased Ca2+ influx [207]. Both events may explain the synaptic dysfunction and neuronal vulnerability in PD. L-type Ca2+ channels are also implicated in PD development, as administration of L-type Ca2+ channel blockers (i.e., isradipine, nimodipine) in animal models and PD patients, reduced death risk and ameliorated disease manifestations [208,209,210,211]. Finally, aSyn oligomers can inhibit α4β2 nicotinic acetylcholine receptors of dopaminergic neurons, thus leading to cholinergic signaling deficits [212]. In summary, aSyn seems to regulate neuronal toxicity and survival via the formation of channels or pores in the plasma membrane or via its interaction with other channels or receptors crucial for the proper neuronal activity (Figure 1).

2.5. Dopamine Metabolism

Soluble aSyn has been proposed to interact with the dopamine transporter (DAT) and decrease its amount on the plasma membrane, thus regulating the dopamine re-uptake from the synapse and protect neuronal cells from excessive dopamine toxicity [213,214]. Contrariwise, aSyn aggregation triggers DAT recruitment to the plasma membrane that results in massive entry of dopamine and production of reactive oxygen species (ROS) in neurons [215]. It is obvious that aSyn-mediated modulation of DAT activity is crucial for neuronal functioning via a balanced dopaminergic neurotransmission. Moreover, the regulation of dopamine storage is provided by an interaction of aSyn with the vesicular monoamine transporter 2 (VMAT2), which is responsible for the packaging of monoamine transmitters into synaptic vesicles [216]. It has been reported that increased levels of aSyn lead to VMAT2 inhibition and dopamine dysregulation that results in pathological events [217]. In addition, aSyn regulates dopamine biosynthesis, via reducing the activity or the phosphorylation status of TH, the rate-limiting enzyme in catecholamine synthesis [218,219,220,221,222,223]. In agreement, enhanced expression or phosphorylation and subsequent aggregation of aSyn alter TH activity and evoke an imbalance in dopamine synthesis, thus leading to neurotoxicity [132,224,225,226]. In vivo evidence further support the role of aSyn in dopamine metabolism, since the absence of aSyn caused decreased reuptake of dopamine, low levels of TH and DAT in the mouse striatum and reduced number of dopaminergic cells in the substantia nigra of aSyn KO mice [227,228,229].

2.6. Interaction with Mitochondria and ER

alpha-Synuclein displays a remarkable conformational flexibility upon macromolecular interactions and can associate with mitochondrial membranes, thus altering mitochondrial function [230,231,232] (Figure 2). There are reports suggesting that aSyn is a physiological regulator of mitochondrial activity [233,234,235], whereas others support the opposite [236,237,238]. Such discrepancies could be attributed to the different synuclein models utilized in each study, taking into account that brain homeostasis is a complex process and in vivo studies recapitulate better the interplay between the various brain components, compared to the isolated in vitro cellular setup. A bidirectional interaction between aSyn aggregation and mitochondrial dysfunction has been implicated in PD pathogenesis. In particular, increased levels of aSyn can lead to mitochondrial dysfunction [239,240,241,242,243,244], whereas, conversely, impairment of mitochondrial activity may accelerate aSyn pathology [245,246,247,248]; however, the precise underlying mechanisms remain to be elucidated. Both WT and mutant aSyn have been shown to interact with mitochondrial elements, altering both mitochondria morphology and function. Specifically, soluble pre-fibrillar aSyn oligomers seem to be responsible for complex I dysfunction, loss of membrane potential, disrupted Ca2+ homeostasis, enhanced cytochrome c release and ROS production, thus leading to neuronal demise [240,249,250,251,252].
Experiments in various animal models of a-synucleinopathy have revealed mitochondrial abnormalities, DNA damage and neuronal degeneration in PD-affected brain regions [244,253,254]. Moreover, in vitro and in vivo experiments have shown that aSyn inhibits mitochondrial fusion and triggers mitochondrial fragmentation [231,255]. Di Maio and colleagues have proposed that certain post-translationally modified aSyn conformations (soluble oligomers, dopamine-modified and S129E phosphorylation mimic) lead to impaired mitochondrial function via binding to TOM20 (translocase of the outer membrane receptor) and inhibiting mitochondrial protein import [239].
Nonetheless, there is evidence suggesting an impairment of mitochondrial function upstream of aSyn pathology. Experiments using the pesticides rotenone and paraquat have shown that dysregulation of mitochondrial function leads to nigrostriatal dopaminergic loss and formation of LB-like inclusions, positively stained with anti-aSyn antibodies and thioflavine S, thus resembling PD features [246,247,256,257,258]. Similarly, incubation of WT aSyn-overexpressing COS-7 cells with mitochondrial inhibitors resulted in the disappearance of the aSyn aggregates formed upon rotenone or oligomycin treatment [259]. A plethora of studies that utilize the mitochondrial neurotoxin MPTP to induce PD-like pathology in animals, further suggest that mitochondria impairment is a key player in disease development [245,248,260,261,262,263,264]. Genetic studies further support the hypothesis of aSyn accumulation as a secondary event following mitochondrial malfunction. Specifically, mutations in ATP13A2 (ATPase cation transporting 13A2), encoding for the lysosomal type 5 P-type ATPase, were shown to result in dysregulation in mitochondrial depolarization and ATP metabolism leading to mitochondrial fragmentation and subsequent cell death [265,266].
Apart from its implication in mitochondrial failure, aSyn has been also reported to play a biological role in the association of mitochondria with the endoplasmic reticulum (ER) Ca2+ homeostasis. It has been demonstrated that aSyn favors the Ca2+ transfer from ER to mitochondria, as a result of the communication the two organelles, probably due to the fact that aSyn can act as a “bridge” via its C terminus [267]. Later studies further supported the physiological localization of aSyn in mitochondria-associated ER membranes (MAM), stabilizing their interaction, which was perturbed upon aSyn aggregation and its subsequent redistribution [268,269]. Interestingly, the familial PD-linked A53T and A30P aSyn point mutations resulted in their weakened interaction with MAM, which affected MAM function and mitochondrial integrity [269].
The association of aSyn with mitochondria was further corroborated by findings indicating interactions between both monomeric and oligomeric aSyn with the Ca2+ transporting voltage-dependent anion channel 1 (VDAC1) [270,271,272,273]. Importantly, VDAC1 has been detected on the MAM of ER mediating the communication between the two organelles, regulating Ca2+ homeostasis [274,275,276]. Moreover, VDAC levels have been found decreased in nigral neurons of PD brains, where pathological aSyn inclusions had been formed [277]. Additionally, VDAC has been proposed to be a component of the mitochondrial permeability transition pore, the opening of which has been shown to be affected by aSyn overexpression and oligomerization [230,278]. In vivo experiments on transgenic mice overexpressing the human A53T aSyn further supported the role of permeability transition pore activity modulation on the mitochondrial dysfunction during PD pathogenesis [279].

2.7. Unfolded Protein Response, Regulation of ER/Golgi Trafficking and Ca2+ Homeostasis

The ER is a continuous membrane system mainly responsible for the production and processing of lipids and proteins, as well as Ca2+ homeostasis. In case of impaired protein folding (ER stress), cells activate a group of signal transduction pathways, known as the unfolded protein response (UPR). It has been previously shown that aSyn overexpression in PD patients leads to UPR and contributes to the molecular pathogenesis of the disease [280]. The ER chaperone glucose regulated protein 78 (GRP78/BIP) has a crucial role on ER stress regulation due to its ability to control the activation of transmembrane ER stress sensors (IRE1, PERK and ATF6) [281]. Disassociation of GRP78 from IRE1 and PERK results in stress signaling, finally leading to altered ER homeostasis [282]. aSyn associates with GRP78/BIP under physiological or pathological conditions, thus inducing UPR and leading to dopaminergic cell death [45,283]. Strikingly, Ser129 phosphorylated and aggregated aSyn was found in ER microsomes of A53T transgenic mice and more importantly, administration of the UPR inhibitor salubrinal, effectively attenuated disease manifestations in this PD-mouse model [284,285]. It is worth mentioning that GRP78/BiP levels were found elevated in DLB and PD brains in an aSyn burden-dependent manner [286]. In addition, the protein levels of various ER chaperones were found elevated in a-synucleinopathy models, co-localized with aSyn positive inclusions, suggesting that aggregated aSyn could potentially be implicated in UPR regulation in disease progression [284,287,288,289,290,291,292,293].
Proteins synthesized in the ER, are packaged into vesicles and directed to Golgi apparatus for subsequent modifications. One of the first pathological roles attributed to aSyn is the blockade of the vesicular transport from ER to Golgi by antagonizing ER/Golgi SNAREs [294,295,296]. Towards the same direction, aSyn can also disrupt the intra-Golgi and post-Golgi secretory trafficking, via an abnormal interaction with several Rab-family proteins of the intracellular endocytic pathway [294,296,297,298,299]. Additionally, aSyn can also impair the ionic transport and membrane trafficking, resulting in Golgi fragmentation and subsequent cytotoxicity [300,301,302].
Another significant role of aSyn on ER and Golgi function is the regulation of Ca2+ homeostasis via its binding on specific channels or pumps localized in these organelles (Figure 2). Specifically, proximity ligation assay experiments demonstrated that soluble and insoluble aSyn aggregates, but not monomers, interact with the ER Ca2+-ATPase SERCA, resulting in decreased cytosolic Ca2+ that disrupts the physiological cell function and leads to neuronal cell death [303]. Moreover, administration of the SERCA inhibitor cyclopiazonic acid restored cytosolic Ca2+ levels and protected neurons against the aggregated aSyn-dependent cell death [303]. In support to these results, aggregated aSyn bound on SERCA pump was detected in LBs and GCIs of PD and MSA brains, respectively [303]. Furthermore, PMR1, a Ca2+-transporting ATPase 1 pump regulating the levels of Ca2+ and Mn+2 ions in the Golgi [304], has been proposed to be a mediator of aSyn-dependent cytotoxicity. Specifically, in various PD models (yeast, flies and nematodes), PMR1 pump has been linked to aSyn pathology via a Ca2+-dependent mechanism, where aSyn accumulation elevated cytosolic Ca2+ levels and increased cell death. Interestingly, upon PMR1 deletion, the disease-associated characteristics were abolished, further suggesting the relevance of this pump to aSyn pathology [305,306].

2.8. a-Synuclein in the Nucleus

The name aSyn was given to the protein due to its localization in the nucleus and presynaptic nerve terminals [37]. Nuclear aSyn was detected in neurons of various brain regions of rodents and was reported to interact with histones, underlying PD pathology [307,308,309], even though a single study declares that the nuclear staining of aSyn is attributed to the non-specific signal of some antibodies that probably recognize unknown antigens in neuronal nuclei [310]. It has been proposed that aSyn is responsible for epigenetic dysregulation via inhibition of histone acetylation or reduced DNA methylation, thus favoring neuronal degeneration, whereas others suggest that nuclear aSyn regulates cell cycle rate exhibiting cell toxicity [311,312,313]. Importantly, histone deacetylase (HDAC) inhibitors attenuated aSyn toxicity and provided neuroprotection in both cell culture and transgenic Drosophila models [311,314].
Experiments in SH-SY5Y cells revealed that nuclear translocation of aSyn is regulated by calreticulin and Ca2+, following treatment with retinoic acid and modulates the expression of PD-linked genes such as ATP13A2 and PINK1 (PTEN-induced kinase1) [315]. Interestingly, phosphorylated aSyn at Ser129 was found accumulated in the nucleus of HEK293E-aSyn overexpressing cells and in various brain regions of transgenic (Thy1)-[A30P] aSyn mice [316]. Further experiments in H4 cells expressing various aSyn proteins verified that nuclear localization of aSyn depends on its phosphorylation at Ser129 [317]. The same group supported a role of DNA-binding and gene expression regulation for aSyn providing an insight into the role of modified aSyn in the nucleus [317]. Furthermore, other post-translational modifications of aSyn, such as sumoylation, seem to be responsible for the translocation of aSyn from the cytoplasm to the nucleus [318]. Although the majority of studies support a neurotoxic role for aSyn nuclear localization, some groups proposed that aSyn in the nucleus displays a protective role against DNA damage, replication stress or impaired nucleo-cytoplasmic transport [319,320,321]. However, the numerous in vitro and in vivo studies demonstrating a neurotoxic role of nuclear aSyn, in contradiction to the limited number of studies supporting a protective role originated mostly from cell lines or yeast, favors the pathological potential of nuclear aSyn.

2.9. Alpha-Synuclein and Protein Degradation Pathways: An Intricate Interplay

A great wealth of data focuses on the complicated relationship between aSyn clearance and protein degradation pathways (Figure 2). Both the ubiquitin-proteasome system (UPS) and the autophagy lysosome pathway (ALP) are responsible for aSyn degradation in a manner that depends on cell type, tissue and aSyn conformation state [322,323,324]. Specifically, there are studies demonstrating that aSyn can be degraded by the 26S/20S proteasome via ubiquitin-dependent [325,326] and ubiquitin-independent manner [327,328]. Studies in PC12, HEK293 and primary mesencephalic cells suggested that pharmacological inhibition of the proteasome does not lead to aSyn accumulation [324,329,330]; however, others have shown that soluble aSyn oligomers, but not monomers, are partially cleared via the 26S proteasome [331]. Importantly, it has been proposed that the UPS is responsible for aSyn removal under normal conditions, while in pathological cases the ALP is recruited to clear the increased aSyn burden [332].
Chaperone-mediated autophagy (CMA) is also responsible for the degradation of monomeric or dimeric forms of the protein via the lysosome-associated membrane protein type 2A (LAMP2A), whereas oligomeric aSyn is cleared mainly via macroautophagy [324,333,334]. Lee and colleagues also suggested that the lysosome is responsible for the removal of oligomeric but not fibrillar aSyn and that lysosomal failure results in aSyn accumulation and aggregation and subsequent cell death [335]. Moreover, initial in vivo evidence suggested that increased aSyn protein levels evoked by paraquat treatment were preferably degraded via CMA in dopaminergic neurons, where the levels of LAMP2A and the lysosomal heat shock cognate protein of 70 kDa (HSC70), both essential CMA-components, were found elevated [336]. We have also shown that boosting CMA function via LAMP2A overexpression in cell lines and primary neuronal cultures and in the rat dopaminergic system mitigated aSyn protein levels and related toxicity [337]. Similar neuroprotective effects were obtained upon LAMP2A overexpression in the Drosophila brain [338]. On the contrary, we have also shown that LAMP2A silencing led to endogenous aSyn accumulation in vitro [324] and in vivo [339] and in extensive neurodegeneration of the rat nigrostriatal axis [339]. Decreased levels of LAMP2A and HSC70 were reported in the human substantia nigra and amygdala of PD brains [340], whereas, in a subsequent study, LAMP2A was found to be selectively reduced in association with increased aSyn levels, even in the early stages of PD, thus suggesting a potential dysregulation of CMA-mediated protein degradation prior to substantial aSyn aggregation in PD [341].
However, a bidirectional link between aSyn accumulation and the protein degradation machineries exists and extensive studies have been conducted to elucidate not only the manner of aberrant aSyn degradation in a-synucleinopathies, but also the impact of various aSyn conformations on UPS and ALP function. It has been proposed that overexpression of A30P and A53T mutants, contrarily to WT aSyn, leads to cell death due to proteasomal inhibition [342]. Indeed, overexpression of mutant A53T aSyn resulted in UPS failure by inhibiting the activity of the 20S/26S proteasome, finally leading to aSyn pathological accumulation [343]. Other groups have failed to detect alterations in the proteasomal function of PC12 cells or transgenic mice, following overexpression of WT or mutant (A30P, A53T) aSyn [344]. Moreover, later studies demonstrated that transient overexpression of WT or mutant aSyn, followed by addition of recombinant aSyn oligomers and fibrils in an osteosarcoma cell line, did not result in any disturbance of the proteasomal function [345]. Importantly, studies in human post-mortem PD brains also suggested impaired proteasomal function in the substantia nigra [346,347,348], further supporting a role of UPS malfunction in PD pathogenesis. In addition, total rates of protein degradation declines with aging, thus contributing to the pathogenesis of age-related diseases [349]. Even though human post-mortem studies provide valuable information in regards to etiology and/or disease pathogenesis, the data obtained should be treated with caution, given into account the overall decline in the function of multiple systems with aging. For a-synucleinopathies, we believe that the use of tissue from affected and non-affected (in regards to aSyn pathology and neuronal death) brain areas may provide useful information regarding early or late events leading to neurodegeneration.
Increased aSyn protein burden is reported to impair macroautophagy function as well, via its interaction with Rab1a, an event that subsequently results in the autophagosome-formation-related protein Atg9 mislocalization [350]. Similar results were obtained from cells expressing the PD-linked mutation of the retromer protein VPS35, which is involved in autophagy and is implicated in PD pathogenesis [351]. The three most well studied PD-linked aSyn mutations, E46K, A30P and A53T, have been shown to promote ALP dysfunction, via either impairing autophagosome formation or inhibiting the selective removal of damaged mitochondria through mitophagy [352,353,354]. It has been previously reported that dopamine-modified aSyn inhibits CMA and this could probably shed light into the selective vulnerability of dopaminergic neurons in PD [355]. Further experiments in human iPSC-derived midbrain dopaminergic neurons revealed that disrupted hydrolase trafficking, due to aSyn overexpression, reduces lysosomal function [356]. Similarly, multiple studies suggest that there is a strong relationship between decreased β-glucocerebrosidase (GCase) activity and aSyn accumulation. In particular, heterozygote mutations in GBA1 gene encoding for β-glucocerebrosidase represent a major risk factor for PD development with a-synucleinopathy [357,358,359,360,361,362].

2.10. Alpha-Synuclein in the Extracellular Space

The first indication that aSyn can be secreted arose from the detection of the protein in human CSF and plasma of PD patients, indicating that aSyn can be released into the extracellular space [363,364] and can exert various deleterious effects on neighboring cells. Further studies supported that aSyn can be secreted from neuronal cells, either via vesicles or exosomes [365,366,367]. Extracellular aSyn has been the subject of intensive research in recent years, mainly due to its propensity to spread from neuron to neuron or other glial cells, as discussed in the following sections.
The major hypothesis regarding the onset and spread of aSyn pathology in a-synucleinopathies relies in the protein’s nucleation propensity that leads to the formation of aberrant aSyn species, which then spread to neighboring cells and tissues via various mechanisms. Furthermore, aSyn has been proposed to act as a “prion-like” protein since it was demonstrated that pathogenic aSyn could transfer from diseased neurons of a PD patient to the healthy transplanted ones, fourteen years after the surgical intervention [368]. Similar results were obtained by other groups in both humans and rats [369,370,371,372]. Experiments of PD and DLB patient-derived brain extracts delivered into the brain of mice and non-human primates further validated the transfer of pathological aSyn and the formation of aSyn aggregates within the recipient neurons [373,374]. Moreover, when Pre-Formed Fibrils (PFFs) were used as seeds in both in vitro and in vivo experiments, the endogenous neuronal aSyn was recruited into the formation of highly insoluble aggregates [104,375,376,377,378,379].
Various mechanisms have been proposed for aSyn spread throughout the nervous system, following its release from neurons where the protein is normally expressed. Candidate mechanisms include aSyn secretion via vesicles, exosomes or even naked protein [364,365,366,380,381,382,383] and its uptake from the cells via conventional endocytosis [384,385], passive diffusion [386], tunneling nanotubes [387], membrane penetration [195,388,389] or receptor-mediated internalization [206,390,391]. Once taken-up by recipient cells, the exogenous aSyn has been shown to trigger the endogenous aSyn accumulation via an unknown mechanism [392,393,394,395]. However, according to the prevailing hypothesis, upon the cell-internalization of aberrant aSyn conformations (oligomers or fibrils), these serve as a template for the recruitment of the endogenous monomeric aSyn into the formation of insoluble aggregates [373,375,376,377,396,397,398]. The prevalently unfolded or alpha-helical aSyn is triggered to self-assemble generating fibrils that subsequently deposit as Lewy bodies [399,400,401].
Neuron-to-neuron aSyn transmission occurs following both anterograde and retrograde axonal transport or trans-synaptic pathways [402,403,404]. Several groups have proposed that dysregulation of axonal transport is implicated in aSyn accumulation at the cell body; however, it is not clear whether PD-linked aSyn mutations play a key role in the process per se [403,405,406,407]. Notably, aSyn in its oligomeric form has been shown to interfere with microtubules and kinesin motors, thus disrupting the anterograde transport and similar results were obtained in an aSyn overexpressing mouse model for PD, as well as in patients diagnosed with the disease [408,409,410]. Additionally, it has been suggested that the variety in a-synucleinopathy phenotypes is attributed to the formation of different aSyn “strains” that display “aggressive” characteristics [17,18,411]. As a consequence of their disparate structures, these “strains” have discrete biochemical responses along the different brain regions and cell types, thus explaining the various disease manifestations of a-synucleinopathies [19,20,125,412,413,414].

3. Glia in the CNS: Scavengers of Extracellular aSyn

3.1. Role in Microglia Function and Dysfunction

Microglial cells are the resident phagocytes of the brain, guarding the CNS homeostasis and performing essential role in health and disease. Specifically, apart from exhibiting immunoreactivity as a response to any changes or inflammatory stimulus, they are responsible for the monitoring and pruning of neuronal synapses [415,416,417,418,419]. Disturbance or loss of brain homeostasis “activates” microglial cells, a term used to describe the changes in their shape, gene expression profile and function during their response [420,421,422,423]. A well-regulated immune surveillance of the brain is essential for the proper CNS functioning; however, an excessive and continuous inflammatory response could potentially lead to cellular and tissue damage, tightly linked to the development of various neurodegenerative diseases [424,425,426,427,428]. Enhanced production of pro-inflammatory cytokines, reactive oxygen species (ROS, NO, superoxide) and glutamate has been shown to lead to dopaminergic cell death in PD [429,430,431].
Indeed, in the diseased brain, microglial cells get activated in two states, M1 and M2, depending on the cytokine signaling pathway involved. The classical pro-inflammatory TNF/IFNγ-mediated activation leads to M1 state, whereas the M2 state is subdivided into the M2a “alternative activation” and the M2c “acquired deactivation” states, acquired following the involvement of anti-inflammatory cytokines IL-4 and IL-13 (for M2a) or IL-10 and TGF-β (for M2c) [432]. It has been proposed that in PD a shift from M2 to M1 phenotype is responsible for disease progression; therefore, the scientific interest has been focused on immunomodulatory therapies promoting the neuroprotective M2 type [433].

3.1.1. Physiological Role of aSyn in Microglia Function

Although aSyn has been primarily characterized as a pre-synaptic neuronal protein, several studies have proposed a physiological role of aSyn in microglial function. Microglial cells from mice lacking aSyn (aSyn -/-) displayed reduced phagocytic activity and enhanced secretion of pro-inflammatory cytokines, thus suggesting that aSyn modulates the activation phenotype of the brain immune cells and contributes to the clearance of debris present in the local brain microenvironment [434,435]. On the other hand, transient overexpression of WT, A53T or A30P aSyn in BV2 microglial cells led to a distinct pro-inflammatory cytokine profile in combination with impaired phagocytic activity [436]. Additionally, microglia isolated from BAC transgenic mice overexpressing aSyn, exhibited dysregulation of cytokine release and phagocytosis [437]. Data obtained from iPSC-derived macrophages from PD patients harboring the A53T aSyn mutation and aSyn triplication mutations further support the implication of aSyn in the phagocytic capability of these cells [438]. Expression of aSyn in microglia has been also proposed to promote cell migration via the enhanced expression of the cell-surface glycoprotein CD44 and the matrix metalloproteinase membrane-type 1 (MMP-MT1) [439]. However, the presence of various aSyn species in the environment of microglial cells alters their physiology and behavior leading to neuroinflammation and neurodegeneration.

3.1.2. Microgliosis in a-Synucleinopathies

Microgliosis is the reaction of CNS microglial cells to pathogenic insults and their shift from a resting to the active state [440]. Since the first study demonstrating microgliosis in PD brains [25], a plethora of reports highlights an important role of activated microglia in disease pathogenesis in both humans and animal models (reviewed in [441,442,443,444,445]). Microglial activation has been observed in PD brains by in vivo positron emission tomography (PET) imaging studies [446,447,448], suggesting that microgliosis is an early event that perpetuates during the disease progression. Additional studies have further supported the hypothesis of an early activation of microglia tightly associated to aSyn pathology, in various PD models [192,449,450,451,452,453,454]. However, other reports suggested that microglia respond differently in the various disease stages, in a manner that depends on the affected brain region and the protein burden of aSyn, indicating the existence of immunological diversity among microglia in the diseased brain [449,450,455].
It has been proposed that the neuron-microglia interaction may contribute to the neuroinflammation that characterizes PD, where neurons expressing aSyn activate microglia, which in turn secrete inflammatory factors surrounding the diseased neurons, thus forming a vicious cycle [456] (Figure 3). Likewise, microgliosis has been also reported in MSA, where aSyn is found aggregated mainly within oligodendrocytes [426,457,458,459]. This hypothesis for neuron-microglia communication in synucleinopathies is further supported by findings demonstrating an altered expression profile of various cytokines in the brains of PD patients [460]. Moreover, microglial activation has been shown to induce aSyn-mediated neuronal cell death, in both in vitro and in vivo PD models [461,462,463]. In addition, microglial cells exposed either to cytokines or to PD-derived CSF, displayed alterations in the intracellular aSyn protein levels, suggesting a crucial role of the brain microenvironment for aSyn accumulation in microglia [464,465]. Activated microglia have been also detected in various transgenic animal models overexpressing wild type of mutated aSyn specifically in neurons [192,452,453,454,466,467]. Experimental PD animal models including the use of the MPTP mouse model or nigral injections of recombinant aSyn fibrils or AAVs overexpressing aSyn are characterized by neuroinflammation followed by a significant degeneration of dopaminergic neurons [449,450,468,469,470]. Similarly, further results obtained from the rAAV-driven overexpression of aSyn in the mouse substantia nigra show extensive aSyn-mediated microgliosis primarily in the nigrostriatal axis, accompanied by an increase in the production of pro-inflammatory cytokines [410,471,472].

3.1.3. Activation of Microglia and Clearance of Toxic aSyn Species

Numerous in vitro and in vivo studies have demonstrated that either conditioned medium from aSyn overexpressing cells or aSyn per se (i.e., recombinant monomeric, oligomeric or fibrillar aSyn) robustly activate microglia. In particular, treatment of microglial cells with non-aggregated aSyn was shown to increase phagocytosis and enhance pro-inflammatory cytokines release, NF-kB nuclear translocation and microglial migration [439,468,473,474,475], whereas addition of fibrillar aSyn in BV2 cells was reported to reduce their phagocytic activity [474]. In agreement, incubation of human microglial cell lines or primary microglial cells with monomeric aSyn triggered the release of various pro-inflammatory factors [476,477,478,479]; however, recently, it has been suggested that monomeric, contrarily to oligomeric aSyn, promotes an anti-inflammatory phenotype of microglia [480]. Other groups have found that aggregated aSyn leads to increased TNFa and ROS production, both related to cell toxicity [456,481,482,483].
Microglial activation may depend on the aggregation state of aSyn and microglial cells readily take-up fibrillar aSyn and produce pro-inflammatory cytokines [484]. Furthermore, incubation of microglial cells with conditioned media from neuronal cells or with CSF from PD patients resulted in significant secretion of TNFa, IL1β and ROS [465,473,478]. Importantly, it has been shown that PD-related aSyn mutants are more prone to activate microglia when compared to WT protein [463,479,485]. Moreover, elevated levels of CXCL12 chemokine in both postmortem PD brain tissue and in nigral microglia of transgenic A53T mice further support the aSyn-mediated neuroinflammation [486]. Contrariwise, aSyn-evoked microgliosis in some instances leads to the enhanced expression and release of neuroprotective factors, such as BDNF, probably in an attempt for neuronal repair and survival [481]. Notably, studies in microglia of mice lacking aSyn expression have verified the critical role of aSyn in modulating microglial activation [434,435].
aSyn has been proposed to enter microglial cells via a clathrin-mediated endocytosis and leads to neurotoxicity in a Prostaglandin E2 receptor subtype 2 (EP2)-dependent manner [470,487]. Integrin CD11b has been proposed as a mediator of aSyn-induced activation of microglial cells [488,489,490], whereas other membrane receptors implicated in microglial stimulation in the presence of aSyn are the protease activated receptor 1 (PAR-1), the macrophage antigen 1-receptor (Mac-1), the Fcγ receptors (FcγR), the microglial purinergic receptor P2X7 and the scavenger receptor CD36, as well as the adhesion molecule CD44 and some plasma membrane ion channels [439,463,477,478,483,491,492,493]. Importantly, Toll-like-receptors 2 and 4 (TLR2 and TLR4) are considered crucial modulators of glial responses in a-synucleinopathies, as well as key players in aSyn phagocytosis. These properties have put them in the spotlight as novel therapeutic modulators of neuroinflammation in neurodegenerative diseases [494,495,496,497,498,499,500,501,502].
The internalization of aSyn by microglia triggers various immune response-related cascades, including NF-kB, Nrf2, MHCII and inflammasome. Numerous studies have reported activation of NF-kB pathway upon addition of various aSyn conformations in both rodent and human microglial cell lines [468,469,476,478,485,503]. The nuclear translocation of NF-kB is a result of aSyn interaction with TLR that leads to the Myd 88-mediated activation of IkB kinases [504]. Moreover, oligomeric aSyn has been shown to trigger TLR2 signaling in microglial cells via NF-kB and p38 MAPK activation, which has been previously linked to aSyn-related toxicity [469,505]. Another key player in both aSyn pathogenesis and the neuroimmune system is LRRK2 (Leucine-rich repeat kinase 2), constitutively expressed in neurons and glial cells, mutations of which have been characterized as common risk factors for PD. Significantly, manipulation of LRRK2 expression levels in mouse microglial cells has revealed its role in regulating aSyn degradation [506,507].
Furthermore, various studies have proposed that aggregated aSyn results in IL-1β production by reactive microglia, which in turn involves nod-like receptor protein 3 (NLRP3) inflammasome activation [508,509,510]. Interestingly, the inflammasome related caspase-1 activation is responsible for aSyn truncation and its subsequent pathological accumulation [188]. Apart from the activation of microglial pro-inflammatory transcription factors, such as NFkB, pathological aSyn also affects the antioxidant transcription factor Nrf2 [485,511]. Modifications in the expression levels of Nrf2 have confirmed its role in aSyn clearance and neuronal survival [512,513,514]. Moreover, since microglial cells act as antigen presenting cells in the brain, it has been proposed that upon aSyn internalization by microglia, the protein is presented to T-cells via MHCII, which then infiltrate in the CNS and finally lead to neuronal degeneration present in PD and MSA [22,515,516,517,518,519,520].
Apart from modulating the immune responses in the CNS, microglial cells are the brain’s phagocyting cells clearing cellular debris or any toxic insult. Amongst the various threats, extracellular aSyn has been shown to be effectively processed by activated microglia, in some cases via a DJ1-mediated autophagy [474,521,522,523,524]. Consistent with these results, experiments utilizing primary cells, have demonstrated that aSyn upon its internalization by microglia is targeted to autophagosomes, thus leading to its degradation [493]. TLR4 also seems to play a crucial role in microglial phagocytosis, since TLR4 (-/-) murine microglial cells exhibit impaired aSyn clearance and enhanced neurotoxicity [494,496]. Furthermore, addition of aSyn fibrils in BV2 and primary microglial cells induces autophagy as a rescue mechanism to restore lysosomal damage [525].
Importantly, aging is another key player in the efficient aSyn clearance, since it has been proposed that microglia and monocytes display reduced phagocytic activity with age [526,527]. Regarding the uptake of aSyn by microglia, it has been reported that ganglioside GM1 and lipid rafts, but not clathrin, caveolae and dynamin, mediate monomeric aSyn internalization, whereas aggregated aSyn enters microglia via clathrin- and calnexin-dependent mechanisms [487,528]. Moreover, microglial uptake of neuronally-derived exosome-associated aSyn via macropinocytosis could potentially account for pathological aSyn spreading [529,530,531]. Strikingly, apart from the immunomodulating role of microglia in the CNS, it has been also suggested that resting microglia regulates the cell-to-cell transfer of aSyn in vivo [532]. Therefore, further considering the aforementioned involvement of exosomes in the transmission of aSyn pathology, targeting exosome-release from various cell types of the brain could be a potential therapeutic target against disease progression.

3.2. Astrocytes in a-Synucleinopathies

Astrocytes, the star-shaped cells of the brain, are the most abundant glial cells of the CNS, accounting for at least one third of the brain mass. They have a supportive role to neurons, by maintaining osmotic, energetic and structural tissue homeostasis. In particular, it is suggested that astrocytes regulate neurotransmitter removal from the extracellular space, synaptic transmission, myelination, brain energy metabolism and pH homeostasis, ion balance, cholesterol synthesis, blood–brain barrier (BBB) permeability, cerebral blood flow and glymphatic system function [533,534,535,536,537,538,539,540,541,542,543]. It has also been suggested that astrocytes modulate neuronal synaptic activity via a Ca2+-dependent release of transmitters that have a feedback action on neurons, a process known as “gliotransmission” [544,545,546]. There are two main types of astrocytes in the brain: protoplasmic and fibrous [547]. Protoplasmic astrocytes are highly branched and are mainly located in the gray matter, tightly associated with neuronal cell bodies and synapses [548], whereas fibrous astrocytes have straight and long processes and they are widely distributed in the white matter, contacting nodes of Ranvier [549].
Apart from their crucial role in brain homeostasis and health, astrocytes have been also implicated in the cascade of events underlying neurodegenerative diseases. Specifically, astrocytes in PD brains have been reported to migrate and become reactive and have been classified in two categories depending on their neuroprotective or neurotoxic effects: harmful A1 astrocytes and protective A2 astrocytes [24]. Activated microglial cells induce A1-astrocyte reactivity, leading to neuronal and oligodendroglial cell death and subsequent synaptic impairment [24]. They have been also shown to secrete inflammatory cytokines, partially as a response to the increased aSyn protein load, thus contributing to PD progression and neurodegeneration [550], although there are studies reporting a neuroprotective role of astrocytes in a-synucleinopathies, modulating the levels of accumulated aSyn and protecting neurons against oxidative stress [551,552].

3.2.1. Astrocytes in PD: Friend or Foe?

Up to date, there is little evidence regarding the expression levels [553] and the physiological role of aSyn in astrocytes, such as the implication of aSyn in astrocytic fatty acid metabolism [90]. It is also suggested that cultured human astrocytes express low levels of aSyn and various inflammatory cytokines or cell stress enhance aSyn production [554,555]. Yet, numerous studies have focused on the role of astrocytes in the modulation of aSyn levels and the regulation of immune responses in neurodegeneration. Interestingly, aSyn-positive inclusions have been detected in astrocytes in various regions of PD and DLB brains [13,14,15,556,557,558,559,560,561]. According to the prevailing hypothesis, the accumulated aSyn detected in astrocytes origins from the neighboring diseased neurons, which upon its release is subsequently internalized by astrocytes, probably as a mechanism of clearance and neuroprotection [562] (Figure 3). However, the responses of astrocytes in aSyn-related pathogenesis have been characterized as a “double-edged sword”, due to the controversial findings regarding their role in neurotoxicity or neuroprotection [563].
The neuron-to-astrocyte aSyn transmission has been extensively studied and is considered as the prime suspect for the detection of aSyn immunoreactivity within astrocytes in a-synucleinopathies. A seminal study utilizing primary astroglial cultures treated with conditioned media from differentiated SH-SY5Y neuronal cells showed that internalization of neuronally-derived aSyn by astrocytes occurs via endocytosis and results in the formation of proteinase-K resistant aSyn inclusions within astrocytes [550]. Similar results were obtained in transgenic mice overexpressing human aSyn under the neuronal promoter PDGFβ, where aSyn-positive inclusion bodies were observed in both neurons and astrocytes [550]. Furthermore, co-culture of primary astrocytes with SH-SY5Y cells verified the formation of LB-like inclusions positive for aSyn staining [564]. The detection of aSyn within the astrocytic endosomal/lysosomal compartment led to the hypothesis that astrocytes normally internalize neuronally-derived aSyn as a neuroprotective mechanism; however, prolonged exposure to pathological aSyn species may lead to impaired glial lysosomal function and, thus, astrocytic aSyn accumulation [550,565]. In agreement, overexpression of PD-linked aSyn mutants (A30P and A53T) in actrocyte cell lines resulted in impaired autophagic function, mitochondrial failure and cellular apoptosis, thus preventing the astrocyte-mediated neuroprotection [566]. In support to a protective role of astrocytes during PD, it has been reported that neurotrophin-immunoreactive (BDNF and NT-3) astrocytes surrounding degenerating nigral neurons in the brains of idiopathic PD patients may indicate a beneficial role of glial cells against neuronal failure [567]. In addition, elevated levels of the glutathione peroxidase-GPx in astrocytes of the substantia nigra of PD patients, further demonstrates their role against neuropathology [568,569]. Strikingly, overexpression of the Nrf2 transcription factor selectively in astrocytes in the haSyn A53T transgenic mouse model provided protection against aSyn-related toxicity by enhancing its degradation via the ALP [513]. In a recently published study, it was suggested that astrocytes internalize aSyn aggregates and effectively degrade them via proteasomal and autophagic pathways, thus protecting dopaminergic neurons against the aberrant effects of toxic aSyn species [570].
In vitro and ex vivo studies proposed that aSyn can efficiently transmit not only from neurons to astrocytes, but also between astrocytes and translocate to the lysosomes of the recipient cells [571]. However, differently from neurons, astrocytes are able to efficiently degrade fibrillar aSyn, suggesting an active role for these cells in clearing aSyn deposits [571]. It is worth-mentioning that the transfer of aSyn from astrocytes to neurons was reported to a lesser extent, even though another study suggested that aSyn deriving from SNCA-flag tagged PD astrocytes was effectively transferred to the co-cultured surrounding control neurons, which then displayed signs of degeneration [572]. Strikingly, iPSC-derived astrocytes from PD neurons displayed impaired CMA and macroautophagy, which could potentially account for the observed aSyn accumulation, highlighting the implication of astroglial-mediated proteolysis in the spread of aSyn pathology [572]. On the contrary, when astrocytes derived from healthy controls were co-cultured with PD neurons, the former cells absorbed the neuronal aSyn from the medium, thus indicating a potential neuroprotective effect of astrocytes [572].
In another aSyn-overexpressing neuron-astrocyte co-culture, addition of aSyn fibrils led to the formation of aSyn aggregates in both cell types, probably recapitulating the events occurring during PD pathogenesis [573]. It has been also suggested that aSyn enters astrocytes more efficiently than neurons via endocytosis and that aSyn transmission from astroglial to neuronal cells leads to neurotoxicity and cell death [574]. Indeed, the uptake of aSyn by astrocytes has been reported as an early-in-time event, since it was detected in the cytoplasm of the cells within 30 min following aSyn application [574]. The spread of aSyn from neurons to astroglial cells has also been demonstrated in rodent transgenic animal models overexpressing human aSyn in neuronal cells, by the detection of aSyn-positive inclusions within the cytoplasm of these glial cells [575,576]. Furthermore, hippocampal delivery of aSyn fibrils in the brain of M83 A53T Tg mice resulted in the formation of pSer129-positive aSyn inclusions within both astrocytes and microglia, four months post-injection [397],
Various mechanisms have been proposed to underlie aSyn transfer between neurons and astrocytes. For example, it is known that astrocytes, upon oxidative stress, form tunneling nanotubes (TNTs) in order to connect with other non-stressed cells [577]. It has been proposed that various stressors, such as aberrant aSyn, lysosomal dysfunction or mitochondrial failure could trigger TNT formation in astrocytes and enhance aSyn spreading [387,578]. Another mechanism for intercellular communication is via exosomes. Although extensive work has been done regarding the role of exosomes in neuronal aSyn transmission and disease pathology [366,579,580,581,582], few studies have proposed a vesicular-mediated transfer of neuroprotective molecules from astrocytes to neurons [583,584] or an exosome-related aSyn spread from neurons to astrocytes [585,586].
Upon aSyn transmission to astrocytes, the latter produce multiple pro-inflammatory cytokines (IL-1α, IL-1β, IL-6, IL-18) and chemokines (CC-, CXC- and CXCL-type) as a response [550,587]. It has been suggested that the pro-inflammatory response of astrocytes to aSyn depends on TLR4 [496,588]. Apart from cytokine release, aSyn leads to Ca2+ flux and oxidative stress upon its entry in astroglial cells, presumably leading to neurotoxicity [589,590,591]. Moreover, in vitro and in vivo experiments have shown that pathological aSyn triggers microglial activation, followed by the induction of reactive A1 astrocytes, finally leading to neurodegeneration [592]. It has been also reported that astrocytes over-expressing hA53T aSyn displayed impaired functions, including glutamate uptake and BBB regulation, resulting in paralysis in transgenic mice [593]. Additionally, aSyn-treated astrocytes have been proposed to produce reduced levels of cholesterol, whereas in parallel they display enhanced GFAP expression, indicative of astrocyte reactivity [594]. Significantly, treatment of astrocytes with various aSyn species (monomeric, oligomeric, fibrillar) induced astrocyte activation and secretion of TNF-α and IL-1β, the expression of which seemed to depend on aSyn species, leading to subsequent non-cell autonomous neuronal degeneration [595]. Other studies revealed that elevated expression levels of myeloperoxidase and enhanced IFN-γ signaling could mediate the astrocytic-activation and inflammation observed in PD brains [21,596,597,598].
alpha-Synuclein has been reported to trigger the opening of connexin 43 (Cx43) hemichannels and pannexin-1 (Panx1) channels in mouse cortical astrocytes, leading to alterations in [Ca2+]i levels, production of nitric oxide (NO), enhanced purinergic and glutamatergic signaling, altered mitochondrial morphology and reduced astrocyte survival [599]. Another recently published study suggests that vesicle-associated aSyn, deriving from erythrocytes, effectively crosses the BBB and accumulates within astrocytes, impairing glutamate uptake, probably due to interactions of oligomeric aSyn with excitatory amino acid transporter 2 (EAAT2) [600]. Furthermore, astrocytes overexpressing mutant A35T and A30P aSyn triggered ER stress and damaged the Golgi apparatus, finally leading to apoptotic cell death [601]. Remarkably, co-culture of primary astrocytes overexpressing mutant aSyn with neuronal cells, inhibited neurite outgrowth, probably due to reduced GDNF secretion [601]. Finally, Cy3-labeled aSyn oligomers were internalized by glial cells, primarily astrocytes, which then started to degrade the oligomers via the ALP [602].
In addition to the well-established role of microglia in the activation of astrocytes, astrocytes themselves have been also reported to control microglial activation and microglia-induced neuroinflammation [603,604], thus unraveling an astrocyte-microglia intimate crosstalk (Figure 4). For example, in response to pathological aSyn insult, astrocytes can acquire a pro-inflammatory phenotype that can lead to neuronal death, independent of microglia. Given that astrocytes produce pro-inflammatory cytokines and chemokines as a response to various stimuli, it has been proposed that such astrocytes may mediate the microglial activation detected in aSyn-related brain diseases [605,606]. Various WT or mutant aSyn conformations have been shown to trigger the up-regulation of pro-inflammatory modulators in astrocytes, such as ICAM-1, IL-6 and TNF-α, leading to microglial activation, neuroinflammation and neurotoxic events during PD progression [496,587,607]. Specifically, transgenic mice inducibly overexpressing the PD-related A53T mutant aSyn selectively in astrocytes exhibited reactive astrogliosis accompanied by increased inflammatory responses and microglial activation in brain regions with significant neuronal loss [593]. Moreover, the detection of ICAM-1 positive reactive astrocytes surrounding brain areas with severe neuronal loss in PD brains or in the MPTP mouse PD model, indicates a sustained inflammatory process mainly triggered by astroglial cells, which is responsible for the consequent degeneration of dopaminergic neurons [608]. Finally, similar to the aSyn prion-like mode of action, the neurotoxic reactive astrocyte polarization has been recently proposed to occur during the CNS prion disease, where microglial cells seem to retain a neuroprotective phenotype against the inflammatory astrocytic responses [609]. All the above observations highly cement the contribution of astroglial reactivity in the pathogenesis of PD and related neuronal a-synucleinopathies.

3.2.2. Implication of astrocytes in MSA pathology

Apart from PD and DLB, astroglial activation is also present in MSA and it seems to play a role in both disease initiation and progression. Extensive astrogliosis has been reported in various brain regions of MSA patients, in some cases accompanied by aSyn-positive inclusions within astrocytes [16,610,611,612,613]. In agreement, accumulation of abnormally phosphorylated and aggregated aSyn was present within astrocytes of MSA patients [614]. Interestingly, aSyn-positive doughnut-shaped inclusions were detected in radial processes of Bergmann glia (unipolar protoplasmic astrocytes in the cerebellar cortex) of various a-synucleinopathy brains, including MSA [558]. Although astroglial aSyn accumulation has been demonstrated in the brains of various MSA transgenic mouse models, finally leading to astrogliosis that accompanies neurodegeneration [615,616,617,618], Song et al. suggested that subcortical astrocytes in MSA did not display aSyn accumulation, in contrast to PD [14]. However, up-to-date the data on the precise role of astrocytes in MSA are scarce, reinforcing the necessity for further studies to elucidate the contribution of astroglial activation in MSA pathogenesis.

3.3. Alpha-Synuclein in Oligodendrocytes: The Pathologic Hallmark of MSA, A Unique Oligodendrogliopathy

Oligodendrocytes are responsible for the production of myelin [619] that surrounds the neuronal processes, mediates the transmission of electric signals between neurons and provides a neurotrophic support [620,621,622,623]. There are two main types of oligodendrocytes, the myelinating and the non-myelinating ones, that concentrate in white and grey matter, respectively [624]. Many neurodegenerative diseases occur due to either oligodendroglial death or damage to the myelin sheathes they produce leading to subsequent neuronal demise [622,625,626]. Oligodendrocytes have been proposed to participate in late PD and DLB progression, rather than in disease initiation [442,627]. This secondary involvement of oligodendrocytes in neuronal a-synucleinopathies was further supported by the detection of aSyn-positive inclusions within non-myelinating oligodendrocytes of PD and DLB brains [15,628], as well as by the presence of complement-activated oligodendrocytes in the diseased brains [629,630]. Moreover, axonal myelination deficits in neurodegenerative diseases also highlight the involvement of oligodendrocytes in neuropathogenetic events [631,632]. Additionally, oligodendroglial cell loss was reported in the striatum of the MPTP-intoxicated mouse PD model, shortly after MPTP administration [633].
On the other hand, the oligodendroglial aSyn inclusions detected in the brains of MSA patients (GCIs) are the main hallmark of the disease and are considered to play a crucial role in the primary events leading to MSA [626]. The involvement of oligodendrocytes in MSA initiation and progression is indisputable, due to the wide distribution of GCIs along the affected brain areas of MSA patients [611,634,635,636,637]. Apart from GCIs present within oligodendrocytes, other inclusions, such as neuronal cytoplasmic inclusions (NCIs) have been detected in neuronal somata, axons and nucleus in various brain regions, mainly composed of aSyn [638,639,640]. GCIs were first described three decades ago, as multi-shaped oligodendroglial inclusions composed of a 10-nm-sized central core fibrils, positively stained with antibodies against aSyn, surrounded by other aggregated proteins such as αβ-crystallin, ubiquitin, cytoskeletal proteins, chaperones and the microtubule-related proteins TPPP/p25α and tau [11,634,641,642,643]. Comparative analysis of the protein composition of GCIs and LBs revealed that GCIs consist of 11.7% aSyn, 1.9% αβ-crystallin and 2.3% 14-3-3 proteins, whereas LBs are composed of 8.5% aSyn, 2% αβ-crystallin and 1.5% 14-3-3 proteins [644]. It has been also reported that aSyn is a key player in the accumulation of tau and αβ-crystallin within GCIs [645], further highlighting the contribution of aSyn in disease pathogenesis.
Another early hallmark of oligodendroglial pathology in MSA and major component of GCIs is the oligodendroglial-specific phosphoprotein TPPP/p25α, which under physiological conditions has been proposed to mediate the myelination process and colocalize with myelin basic protein (MBP) in normal human brains [646,647,648]. Under pathological conditions, TPPP/p25α is considered to re-locate from the myelin sheaths to the abnormally expanded oligodendroglial somata and to trigger aSyn aggregation in vitro [12,649]. In vitro experiments utilizing TPPP/p25α ectopic overexpression in PC12 cells revealed that TPPP/p25α prevents the fusion of autophagosomes with lysosomes and impairs aSyn degradation, enhancing its secretion via exophagy [650]. Moreover, concurrent overexpression of TPPP/p25α and aSyn in OLN-93 rat oligodendroglial cells led to pSer129 aSyn-dependent microtubule retraction from the processes to the perinuclear space, as well as to cytotoxicity and subsequent cell death via activation of the FAS receptor and caspase-8 [651,652]. Recently published work from our lab revealed a crucial role for TPPP/p25α in the recruitment and seeding of oligodendroglial aSyn and in the formation of aberrant aSyn species within oligodendrocytes [653]. Additionally, the levels of glial cell-derived neurotrophic factor (GDNF) were found significantly decreased in the brains of MBP-haSyn transgenic mice, a mouse model for MSA where human aSyn is specifically overexpressed in oligodendrocytes [654]. Similar results were obtained from brain samples of MSA patients, further supporting that oligodendroglial aSyn accumulation is implicated in the dysregulation of neurotrophic support, oxidative stress and neuroinflammation, thus leading to MSA pathogenesis [654].

Alpha-Synuclein Accumulation in Oligodendrocytes, Propagation and Spread of Pathology

The origin of aSyn detected in oligodendroglial GCIs still remains enigmatic and there are controversial studies in the literature suggesting either the internalization of neuronally-secreted aSyn by oligodendrocytes or an enhanced expression and decreased degradation of oligodendroglial aSyn [655,656,657,658,659,660,661]. The release of aSyn by neuronal cells, partially bound on vesicles or exosomes is well-documented [365,366,380,662,663] and some studies propose that oligodendrocytes can take-up the neuronally-derived aSyn via dynamin GTPase-, clathrin- and dynasore-dependent mechanisms [658,664,665,666]. The neuron-oligodendrocyte communication can also be mediated via exosomes [667], which are characterized as “Trojan horses” of neurodegeneration [668] and they could serve as transporters of pathological disease-related proteins, such as aSyn (Figure 3). Moreover, ectopic expression of the endocytosis regulatory proteins Rab5 and Rabaptin-5 in oligodendrocytes of MSA brains may account for the elevated levels of aSyn within oligodendrocytes, probably via enhanced endocytotic activity [669].
In vitro and in vivo experiments revealed the ability of oligodendrocytes to take up exogenously added recombinant or neuronally-derived aSyn and incorporate it into intracellular GCI-like aggregates [658,660,664,665,666]. In a recently published study, mature human oligodendrocytes generated from neural stem cells had the ability to internalize neuronally-derived aSyn and form proteinaceous inclusions, thus further supporting the existing theory for the origin of MSA-related aSyn. Significantly, it has been shown that once neuronal aSyn is taken up by oligodendrocytes, it accumulates and gains GCI-like characteristics rather determined by the oligodendroglial milieu [17].
Over the last years, the prion hypothesis has gained a lot of attention regarding the spread of pathological aSyn in the context of both PD and MSA. Specifically, it has been reported that inoculation of transgenic mice overexpressing human A53T aSyn with MSA brain homogenates resulted in CNS dysfunction, whereas, strikingly, the PD brain-derived material did not evoke similar effects [670,671,672]. Similarly, intrastriatal injections of MSA homogenates in the brains of Tg(SNCA)1Nbm/J mice (out for mouse aSyn and overexpressing the human protein) resulted in the detection of hyper-phosphorylated aSyn-positive inclusions in various brain regions [673]. Finally, treatment of HEK293T cells stably expressing fluorescently-tagged aSyn with healthy, PD or MSA brain-derived extracts highlighted that only the MSA-added material was capable of inducing aSyn aggregation [674].
There are other possible scenarios that have been proposed to explain aSyn seeding and propagation in MSA brains, based on oligodendrocyte-to-oligodendrocyte communications. Specifically, it is possible that aSyn is taken up by oligodendroglial progenitor cells prior to their maturation, probably impairing the myelination process [675]. Finally, another scenario suggests that, in MSA pathology, oligodendrocytes adopt unknown cellular uptake mechanisms for aSyn internalization and subsequent propagation [675]; however, the precise mechanisms underlying aSyn transfer to oligodendrocytes still remains unknown. The gap junction protein connexin-32 (Cx32) has been also implicated in the uptake of oligomeric aSyn by both neurons and oligodendrocytes [676] and Cx32 protein levels were found elevated in animal PD and MSA models, thus suggesting an interaction between Cx32 expression and aSyn cellular uptake [676]. Contrariwise, others and we have suggested that the endogenous aSyn, expressed even at minute amounts, plays a pivotal role in the accumulation of pathological aSyn within oligodendrocytes and the subsequent GCI-like formation [653,677].
It is interesting to note that oligodendroglial and neuronal aSyn accumulation has been reported to occur in different time points and in particular that aSyn aggregation requires several months to progress within oligodendrocytes, upon synthetic haSyn-PFFs delivery into the brain of WT mice [678]. Moreover, in vitro aSyn overexpression in oligodendrocytes resulted in delayed maturation of oligodendrocyte progenitor cells and impaired myelin-gene expression and myelination deficits [679,680,681], whereas in another study aSyn-positive inclusions were mainly detected in BCAS1-expressing (breast carcinoma amplified sequence 1) immature oligodendrocytes of MSA brains [682]. The above observations insinuate that the oligodendroglial maturation and aSyn-aggregate formation are closely linked and may provide information regarding pathogenic events in MSA.
Regarding the hypothesis of impaired aSyn degradation in the context of MSA, both the UPS and the ALP have been proposed to contribute to the accumulation and aggregation of aSyn within oligodendrocytes. The detection of LC3-positive signal or other autophagy-related proteins, such as ubiquitin and p62 in GCIs points a role of the ALP in MSA pathogenesis [659,683,684,685,686,687]. It has also been suggested that AMBRA1, an upstream protein regulator of autophagy and UCH-L1, a deubiquitylating enzyme, are implicated in neurodegenerative diseases with oligodendroglia pathology [688,689]. The role of autophagic dysregulation along with mitochondrial impairment in aSyn aggregation was also studied in primary oligodendroglial cultures and in the OLN-t40 oligodendroglial cell line [690]. Moreover, neurosin (kallikrein 6) has been proven an effective serine protease in clearing aSyn from oligodendrocytes both in vitro and in vivo [691,692,693]. In addition, treatment of Tg haSyn-PLP mice, a well-established MSA mouse model, with the proteasome inhibitor I for 12 weeks, resulted in enhanced accumulation of both human and endogenous mouse aSyn within the cytoplasm of oligodendrocytes, thus highlighting the role of UPS in aSyn degradation [694]. Finally, several in vitro studies have proposed that aSyn aggregation is stimulated by heparin and heparan sulfate [391,695,696], linear polysaccharides (glycosaminoglycans) found on the cell membrane and in the extracellular matrix [697,698]. Heparan sulfate has been suggested to mediate aSyn fibril uptake by oligodendrocytes via binding to the plasma membrane [391,699], whereas other have proposed that heparin and heparan sulfate proteoglycans (HSPGs) are responsible for aSyn fibrillation [695,696,700,701,702].
Apart from aSyn toxicity per se, the overexpression of aSyn in oligodendrocytes can lead to oligodendroglial cell death and subsequent neuronal loss via a plethora of mechanisms. Specifically, aSyn-overexpressing oligodendrocytes are more susceptible to oxidative stress and various cytokine actions [591,703], or display impaired adhesion properties [704]. Furthermore, it has been reported that animal MSA models exhibit myelin loss and impaired mitochondrial function, accompanied by severe neurodegeneration in various brain regions [615,617,618,705,706].

4. Conclusions

It has been almost 25 years since the discovery that the Lewy pathology in PD and DLB neurons is immunoreactive for aSyn and, at the same time, neuronal aSyn accumulates in glial inclusions within MSA oligodendrocytes. Even though a plethora of studies focuses on the role of aSyn in neuronal physiology and pathology, increasing amount of data reinforces the contribution of non-cell autonomous neuron-glial interactions in the initiation and progression of a-synucleinopathies. Microglia and astrocytes form the brain’s defense system against neurotoxic insults, become activated and release pro-inflammatory factors. However, uncontrolled activation results in chronic microgliosis and astrogliosis that may be detrimental and lead to neurodegeneration. Even more, the deposition of aSyn in oligodendrocytes impairs their myelinating activity and reduces neuronal trophic support, events that eventually result in neuronal demise. Remarkably, neurons, microglia, astrocytes and oligodendrocytes are all able to take up and clear extracellular aSyn; however, glial cells appear to be the most potent scavengers. The endocytosis of various aSyn species might be conformation-sensitive, cell- and receptor-type specific, adding further complexity in disease management.
Undoubtedly, better understanding of the mechanisms mediating the interaction between neurons and glial cells in a-synucleinopathies may provide insights into neuronal dysfunction and death and may uncover novel disease modifying therapies.

Author Contributions

P.M.; writing—original draft preparation, M.X.; writing—review and final editing. Both authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by an MSA UK Trust grant (2019/MX60185), an MSA Coalition grant (2020-05-001) and an MJFF grant (16887).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

M.X. is supported by an MSA UK Trust grant (2019/MX60185), an MSA Coalition grant (2020-05-001) and an MJFF grant (16887). This research is co-financed by Greece and the European Union (European Social Fund—ESF) through the Operational Programme «Human Resources Development, Education and Lifelong Learning» in the context of the project “Strengthening Human Resources Research Potential via Doctorate Research” (MIS-5000432), implemented by the State Scholarships Foundation (ΙΚΥ)», awarded to PM.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Uversky, V.N. A protein-chameleon: Conformational plasticity of alpha-synuclein, a disordered protein involved in neurodegenerative disorders. J. Biomol. Struct. Dyn. 2003, 21, 211–234. [Google Scholar] [CrossRef]
  2. Spillantini, M.G.; Crowther, R.A.; Jakes, R.; Hasegawa, M.; Goedert, M. alpha-Synuclein in filamentous inclusions of Lewy bodies from Parkinson’s disease and dementia with lewy bodies. Proc. Natl. Acad. Sci. USA 1998, 95, 6469–6473. [Google Scholar] [CrossRef] [Green Version]
  3. Polymeropoulos, M.H.; Lavedan, C.; Leroy, E.; Ide, S.E.; Dehejia, A.; Dutra, A.; Pike, B.; Root, H.; Rubenstein, J.; Boyer, R.; et al. Mutation in the alpha-synuclein gene identified in families with Parkinson’s disease. Science 1997, 276, 2045–2047. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Chartier-Harlin, M.C.; Kachergus, J.; Roumier, C.; Mouroux, V.; Douay, X.; Lincoln, S.; Levecque, C.; Larvor, L.; Andrieux, J.; Hulihan, M.; et al. Alpha-synuclein locus duplication as a cause of familial Parkinson’s disease. Lancet 2004, 364, 1167–1169. [Google Scholar] [CrossRef]
  5. Kruger, R.; Kuhn, W.; Muller, T.; Woitalla, D.; Graeber, M.; Kosel, S.; Przuntek, H.; Epplen, J.T.; Schols, L.; Riess, O. Ala30Pro mutation in the gene encoding alpha-synuclein in Parkinson’s disease. Nat. Genet. 1998, 18, 106–108. [Google Scholar] [CrossRef] [PubMed]
  6. Singleton, A.B.; Farrer, M.; Johnson, J.; Singleton, A.; Hague, S.; Kachergus, J.; Hulihan, M.; Peuralinna, T.; Dutra, A.; Nussbaum, R.; et al. alpha-Synuclein locus triplication causes Parkinson’s disease. Science 2003, 302, 841. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Zarranz, J.J.; Alegre, J.; Gomez-Esteban, J.C.; Lezcano, E.; Ros, R.; Ampuero, I.; Vidal, L.; Hoenicka, J.; Rodriguez, O.; Atares, B.; et al. The new mutation, E46K, of alpha-synuclein causes Parkinson and Lewy body dementia. Ann. Neurol. 2004, 55, 164–173. [Google Scholar] [CrossRef]
  8. Pals, P.; Lincoln, S.; Manning, J.; Heckman, M.; Skipper, L.; Hulihan, M.; Van den Broeck, M.; De Pooter, T.; Cras, P.; Crook, J.; et al. alpha-Synuclein promoter confers susceptibility to Parkinson’s disease. Ann. Neurol. 2004, 56, 591–595. [Google Scholar] [CrossRef]
  9. Fuchs, J.; Tichopad, A.; Golub, Y.; Munz, M.; Schweitzer, K.J.; Wolf, B.; Berg, D.; Mueller, J.C.; Gasser, T. Genetic variability in the SNCA gene influences alpha-synuclein levels in the blood and brain. FASEB J. 2008, 22, 1327–1334. [Google Scholar] [CrossRef]
  10. Fanciulli, A.; Stankovic, I.; Krismer, F.; Seppi, K.; Levin, J.; Wenning, G.K. Multiple system atrophy. Int. Rev. Neurobiol. 2019, 149, 137–192. [Google Scholar] [CrossRef]
  11. Spillantini, M.G.; Crowther, R.A.; Jakes, R.; Cairns, N.J.; Lantos, P.L.; Goedert, M. Filamentous alpha-synuclein inclusions link multiple system atrophy with Parkinson’s disease and dementia with Lewy bodies. Neurosci. Lett. 1998, 251, 205–208. [Google Scholar] [CrossRef]
  12. Song, Y.J.; Lundvig, D.M.; Huang, Y.; Gai, W.P.; Blumbergs, P.C.; Hojrup, P.; Otzen, D.; Halliday, G.M.; Jensen, P.H. p25alpha relocalizes in oligodendroglia from myelin to cytoplasmic inclusions in multiple system atrophy. Am. J. Pathol. 2007, 171, 1291–1303. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Braak, H.; Sastre, M.; Del Tredici, K. Development of alpha-synuclein immunoreactive astrocytes in the forebrain parallels stages of intraneuronal pathology in sporadic Parkinson’s disease. Acta Neuropathol. 2007, 114, 231–241. [Google Scholar] [CrossRef] [PubMed]
  14. Song, Y.J.; Halliday, G.M.; Holton, J.L.; Lashley, T.; O’Sullivan, S.S.; McCann, H.; Lees, A.J.; Ozawa, T.; Williams, D.R.; Lockhart, P.J.; et al. Degeneration in different parkinsonian syndromes relates to astrocyte type and astrocyte protein expression. J. Neuropathol. Exp. Neurol. 2009, 68, 1073–1083. [Google Scholar] [CrossRef] [Green Version]
  15. Wakabayashi, K.; Hayashi, S.; Yoshimoto, M.; Kudo, H.; Takahashi, H. NACP/alpha-synuclein-positive filamentous inclusions in astrocytes and oligodendrocytes of Parkinson’s disease brains. Acta Neuropathol. 2000, 99, 14–20. [Google Scholar] [CrossRef]
  16. Wenning, G.K.; Jellinger, K.A. The role of alpha-synuclein in the pathogenesis of multiple system atrophy. Acta Neuropathol. 2005, 109, 129–140. [Google Scholar] [CrossRef]
  17. Peng, C.; Gathagan, R.J.; Covell, D.J.; Medellin, C.; Stieber, A.; Robinson, J.L.; Zhang, B.; Pitkin, R.M.; Olufemi, M.F.; Luk, K.C.; et al. Cellular milieu imparts distinct pathological alpha-synuclein strains in alpha-synucleinopathies. Nature 2018, 557, 558–563. [Google Scholar] [CrossRef]
  18. Bousset, L.; Pieri, L.; Ruiz-Arlandis, G.; Gath, J.; Jensen, P.H.; Habenstein, B.; Madiona, K.; Olieric, V.; Bockmann, A.; Meier, B.H.; et al. Structural and functional characterization of two alpha-synuclein strains. Nat. Commun. 2013, 4, 2575. [Google Scholar] [CrossRef] [Green Version]
  19. Shahnawaz, M.; Mukherjee, A.; Pritzkow, S.; Mendez, N.; Rabadia, P.; Liu, X.; Hu, B.; Schmeichel, A.; Singer, W.; Wu, G.; et al. Discriminating alpha-synuclein strains in Parkinson’s disease and multiple system atrophy. Nature 2020, 578, 273–277. [Google Scholar] [CrossRef]
  20. Candelise, N.; Schmitz, M.; Llorens, F.; Villar-Pique, A.; Cramm, M.; Thom, T.; da Silva Correia, S.M.; da Cunha, J.E.G.; Mobius, W.; Outeiro, T.F.; et al. Seeding variability of different alpha synuclein strains in synucleinopathies. Ann. Neurol. 2019, 85, 691–703. [Google Scholar] [CrossRef] [PubMed]
  21. Hashioka, S.; Klegeris, A.; Schwab, C.; McGeer, P.L. Interferon-gamma-dependent cytotoxic activation of human astrocytes and astrocytoma cells. Neurobiol. Aging 2009, 30, 1924–1935. [Google Scholar] [CrossRef]
  22. Imamura, K.; Hishikawa, N.; Sawada, M.; Nagatsu, T.; Yoshida, M.; Hashizume, Y. Distribution of major histocompatibility complex class II-positive microglia and cytokine profile of Parkinson’s disease brains. Acta Neuropathol. 2003, 106, 518–526. [Google Scholar] [CrossRef] [PubMed]
  23. Doorn, K.J.; Moors, T.; Drukarch, B.; van de Berg, W.; Lucassen, P.J.; van Dam, A.M. Microglial phenotypes and toll-like receptor 2 in the substantia nigra and hippocampus of incidental Lewy body disease cases and Parkinson’s disease patients. Acta Neuropathol. Commun. 2014, 2, 90. [Google Scholar] [CrossRef] [Green Version]
  24. Liddelow, S.A.; Guttenplan, K.A.; Clarke, L.E.; Bennett, F.C.; Bohlen, C.J.; Schirmer, L.; Bennett, M.L.; Munch, A.E.; Chung, W.S.; Peterson, T.C.; et al. Neurotoxic reactive astrocytes are induced by activated microglia. Nature 2017, 541, 481–487. [Google Scholar] [CrossRef]
  25. McGeer, P.L.; Itagaki, S.; Boyes, B.E.; McGeer, E.G. Reactive microglia are positive for HLA-DR in the substantia nigra of Parkinson’s and Alzheimer’s disease brains. Neurology 1988, 38, 1285–1291. [Google Scholar] [CrossRef] [PubMed]
  26. Iwai, A.; Masliah, E.; Yoshimoto, M.; Ge, N.; Flanagan, L.; de Silva, H.A.; Kittel, A.; Saitoh, T. The precursor protein of non-A beta component of Alzheimer’s disease amyloid is a presynaptic protein of the central nervous system. Neuron 1995, 14, 467–475. [Google Scholar] [CrossRef] [Green Version]
  27. Withers, G.S.; George, J.M.; Banker, G.A.; Clayton, D.F. Delayed localization of synelfin (synuclein, NACP) to presynaptic terminals in cultured rat hippocampal neurons. Brain Res. Dev. Brain Res. 1997, 99, 87–94. [Google Scholar] [CrossRef]
  28. Andringa, G.; Du, F.; Chase, T.N.; Bennett, M.C. Mapping of rat brain using the Synuclein-1 monoclonal antibody reveals somatodendritic expression of alpha-synuclein in populations of neurons homologous to those vulnerable to Lewy body formation in human synucleopathies. J. Neuropathol. Exp. Neurol. 2003, 62, 1060–1075. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Barbour, R.; Kling, K.; Anderson, J.P.; Banducci, K.; Cole, T.; Diep, L.; Fox, M.; Goldstein, J.M.; Soriano, F.; Seubert, P.; et al. Red blood cells are the major source of alpha-synuclein in blood. Neurodegener. Dis. 2008, 5, 55–59. [Google Scholar] [CrossRef]
  30. Askanas, V.; Engel, W.K.; Alvarez, R.B.; McFerrin, J.; Broccolini, A. Novel immunolocalization of alpha-synuclein in human muscle of inclusion-body myositis, regenerating and necrotic muscle fibers, and at neuromuscular junctions. J. Neuropathol. Exp. Neurol. 2000, 59, 592–598. [Google Scholar] [CrossRef] [Green Version]
  31. Bottner, M.; Zorenkov, D.; Hellwig, I.; Barrenschee, M.; Harde, J.; Fricke, T.; Deuschl, G.; Egberts, J.H.; Becker, T.; Fritscher-Ravens, A.; et al. Expression pattern and localization of alpha-synuclein in the human enteric nervous system. Neurobiol. Dis. 2012, 48, 474–480. [Google Scholar] [CrossRef]
  32. Aldecoa, I.; Navarro-Otano, J.; Stefanova, N.; Sprenger, F.S.; Seppi, K.; Poewe, W.; Cuatrecasas, M.; Valldeoriola, F.; Gelpi, E.; Tolosa, E. Alpha-synuclein immunoreactivity patterns in the enteric nervous system. Neurosci. Lett. 2015, 602, 145–149. [Google Scholar] [CrossRef]
  33. Li, J.; Henning Jensen, P.; Dahlstrom, A. Differential localization of alpha-, beta- and gamma-synucleins in the rat CNS. Neuroscience 2002, 113, 463–478. [Google Scholar] [CrossRef]
  34. Taguchi, K.; Watanabe, Y.; Tsujimura, A.; Tanaka, M. Expression of alpha-synuclein is regulated in a neuronal cell type-dependent manner. Anat. Sci. Int. 2019, 94, 11–22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Taguchi, K.; Watanabe, Y.; Tsujimura, A.; Tanaka, M. Brain region-dependent differential expression of alpha-synuclein. J. Comp. Neurol. 2016, 524, 1236–1258. [Google Scholar] [CrossRef] [PubMed]
  36. George, J.M.; Jin, H.; Woods, W.S.; Clayton, D.F. Characterization of a novel protein regulated during the critical period for song learning in the zebra finch. Neuron 1995, 15, 361–372. [Google Scholar] [CrossRef] [Green Version]
  37. Maroteaux, L.; Campanelli, J.T.; Scheller, R.H. Synuclein: A neuron-specific protein localized to the nucleus and presynaptic nerve terminal. J. Neurosci. 1988, 8, 2804–2815. [Google Scholar] [CrossRef] [Green Version]
  38. Burre, J.; Sharma, M.; Tsetsenis, T.; Buchman, V.; Etherton, M.R.; Sudhof, T.C. Alpha-synuclein promotes SNARE-complex assembly in vivo and in vitro. Science 2010, 329, 1663–1667. [Google Scholar] [CrossRef] [Green Version]
  39. Zaltieri, M.; Grigoletto, J.; Longhena, F.; Navarria, L.; Favero, G.; Castrezzati, S.; Colivicchi, M.A.; Della Corte, L.; Rezzani, R.; Pizzi, M.; et al. alpha-synuclein and synapsin III cooperatively regulate synaptic function in dopamine neurons. J. Cell Sci. 2015, 128, 2231–2243. [Google Scholar] [CrossRef] [Green Version]
  40. Chen, R.H.C.; Wislet-Gendebien, S.; Samuel, F.; Visanji, N.P.; Zhang, G.; Marsilio, D.; Langman, T.; Fraser, P.E.; Tandon, A. alpha-Synuclein membrane association is regulated by the Rab3a recycling machinery and presynaptic activity. J. Biol. Chem. 2013, 288, 7438–7449. [Google Scholar] [CrossRef] [Green Version]
  41. Greten-Harrison, B.; Polydoro, M.; Morimoto-Tomita, M.; Diao, L.; Williams, A.M.; Nie, E.H.; Makani, S.; Tian, N.; Castillo, P.E.; Buchman, V.L.; et al. alphabetagamma-Synuclein triple knockout mice reveal age-dependent neuronal dysfunction. Proc. Natl. Acad. Sci. USA 2010, 107, 19573–19578. [Google Scholar] [CrossRef] [Green Version]
  42. Burre, J.; Sharma, M.; Sudhof, T.C. alpha-Synuclein assembles into higher-order multimers upon membrane binding to promote SNARE complex formation. Proc. Natl. Acad. Sci. USA 2014, 111, E4274–E4283. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Lautenschlager, J.; Stephens, A.D.; Fusco, G.; Strohl, F.; Curry, N.; Zacharopoulou, M.; Michel, C.H.; Laine, R.; Nespovitaya, N.; Fantham, M.; et al. C-terminal calcium binding of alpha-synuclein modulates synaptic vesicle interaction. Nat. Commun. 2018, 9, 712. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Burre, J.; Sharma, M.; Sudhof, T.C. Definition of a molecular pathway mediating alpha-synuclein neurotoxicity. J. Neurosci. 2015, 35, 5221–5232. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Betzer, C.; Movius, A.J.; Shi, M.; Gai, W.P.; Zhang, J.; Jensen, P.H. Identification of synaptosomal proteins binding to monomeric and oligomeric alpha-synuclein. PLoS ONE 2015, 10, e0116473. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Scott, D.; Roy, S. alpha-Synuclein inhibits intersynaptic vesicle mobility and maintains recycling-pool homeostasis. J. Neurosci. 2012, 32, 10129–10135. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Vargas, K.J.; Makani, S.; Davis, T.; Westphal, C.H.; Castillo, P.E.; Chandra, S.S. Synucleins regulate the kinetics of synaptic vesicle endocytosis. J. Neurosci. 2014, 34, 9364–9376. [Google Scholar] [CrossRef]
  48. Zhang, L.; Zhang, C.; Zhu, Y.; Cai, Q.; Chan, P.; Ueda, K.; Yu, S.; Yang, H. Semi-quantitative analysis of alpha-synuclein in subcellular pools of rat brain neurons: An immunogold electron microscopic study using a C-terminal specific monoclonal antibody. Brain Res. 2008, 1244, 40–52. [Google Scholar] [CrossRef] [PubMed]
  49. Ostrerova, N.; Petrucelli, L.; Farrer, M.; Mehta, N.; Choi, P.; Hardy, J.; Wolozin, B. alpha-Synuclein shares physical and functional homology with 14-3-3 proteins. J. Neurosci. 1999, 19, 5782–5791. [Google Scholar] [CrossRef] [Green Version]
  50. Kim, T.D.; Choi, E.; Rhim, H.; Paik, S.R.; Yang, C.H. Alpha-synuclein has structural and functional similarities to small heat shock proteins. Biochem. Biophys. Res. Commun. 2004, 324, 1352–1359. [Google Scholar] [CrossRef] [PubMed]
  51. Chandra, S.; Gallardo, G.; Fernandez-Chacon, R.; Schluter, O.M.; Sudhof, T.C. Alpha-synuclein cooperates with CSPalpha in preventing neurodegeneration. Cell 2005, 123, 383–396. [Google Scholar] [CrossRef]
  52. Kim, T.D.; Paik, S.R.; Yang, C.H. Structural and functional implications of C-terminal regions of alpha-synuclein. Biochemistry 2002, 41, 13782–13790. [Google Scholar] [CrossRef]
  53. Kim, T.D.; Paik, S.R.; Yang, C.H.; Kim, J. Structural changes in alpha-synuclein affect its chaperone-like activity in vitro. Protein Sci. 2000, 9, 2489–2496. [Google Scholar] [CrossRef] [Green Version]
  54. Park, S.M.; Jung, H.Y.; Kim, T.D.; Park, J.H.; Yang, C.H.; Kim, J. Distinct roles of the N-terminal-binding domain and the C-terminal-solubilizing domain of alpha-synuclein, a molecular chaperone. J. Biol. Chem. 2002, 277, 28512–28520. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Rekas, A.; Ahn, K.J.; Kim, J.; Carver, J.A. The chaperone activity of alpha-synuclein: Utilizing deletion mutants to map its interaction with target proteins. Proteins 2012, 80, 1316–1325. [Google Scholar] [CrossRef]
  56. Srivastava, T.; Raj, R.; Dubey, A.; Kumar, D.; Chaturvedi, R.K.; Sharma, S.K.; Priya, S. Fast kinetics of environmentally induced alpha-synuclein aggregation mediated by structural alteration in NAC region and result in structure dependent cytotoxicity. Sci. Rep. 2020, 10, 18412. [Google Scholar] [CrossRef] [PubMed]
  57. Burre, J.; Vivona, S.; Diao, J.; Sharma, M.; Brunger, A.T.; Sudhof, T.C. Properties of native brain alpha-synuclein. Nature 2013, 498, E4–E6. [Google Scholar] [CrossRef]
  58. Celej, M.S.; Sarroukh, R.; Goormaghtigh, E.; Fidelio, G.D.; Ruysschaert, J.M.; Raussens, V. Toxic prefibrillar alpha-synuclein amyloid oligomers adopt a distinctive antiparallel beta-sheet structure. Biochem. J. 2012, 443, 719–726. [Google Scholar] [CrossRef] [PubMed]
  59. Li, H.T.; Du, H.N.; Tang, L.; Hu, J.; Hu, H.Y. Structural transformation and aggregation of human alpha-synuclein in trifluoroethanol: Non-amyloid component sequence is essential and beta-sheet formation is prerequisite to aggregation. Biopolymers 2002, 64, 221–226. [Google Scholar] [CrossRef] [PubMed]
  60. Chaudhary, H.; Subramaniam, V.; Claessens, M. Direct visualization of model membrane remodeling by alpha-synuclein fibrillization. ChemPhysChem 2017, 18, 1620–1626. [Google Scholar] [CrossRef] [Green Version]
  61. Galvagnion, C.; Brown, J.W.; Ouberai, M.M.; Flagmeier, P.; Vendruscolo, M.; Buell, A.K.; Sparr, E.; Dobson, C.M. Chemical properties of lipids strongly affect the kinetics of the membrane-induced aggregation of alpha-synuclein. Proc. Natl. Acad. Sci. USA 2016, 113, 7065–7070. [Google Scholar] [CrossRef] [Green Version]
  62. Galvagnion, C. The Role of Lipids Interacting with alpha-Synuclein in the Pathogenesis of Parkinson’s Disease. J. Parkinsons Dis. 2017, 7, 433–450. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Narayanan, V.; Scarlata, S. Membrane binding and self-association of alpha-synucleins. Biochemistry 2001, 40, 9927–9934. [Google Scholar] [CrossRef]
  64. Zhu, M.; Fink, A.L. Lipid binding inhibits alpha-synuclein fibril formation. J. Biol. Chem. 2003, 278, 16873–16877. [Google Scholar] [CrossRef] [Green Version]
  65. Lee, H.J.; Choi, C.; Lee, S.J. Membrane-bound alpha-synuclein has a high aggregation propensity and the ability to seed the aggregation of the cytosolic form. J. Biol. Chem. 2002, 277, 671–678. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Cole, N.B.; Murphy, D.D.; Grider, T.; Rueter, S.; Brasaemle, D.; Nussbaum, R.L. Lipid droplet binding and oligomerization properties of the Parkinson’s disease protein alpha-synuclein. J. Biol. Chem. 2002, 277, 6344–6352. [Google Scholar] [CrossRef] [Green Version]
  67. Perni, M.; Galvagnion, C.; Maltsev, A.; Meisl, G.; Muller, M.B.; Challa, P.K.; Kirkegaard, J.B.; Flagmeier, P.; Cohen, S.I.; Cascella, R.; et al. A natural product inhibits the initiation of alpha-synuclein aggregation and suppresses its toxicity. Proc. Natl. Acad. Sci. USA 2017, 114, E1009–E1017. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Fares, M.B.; Ait-Bouziad, N.; Dikiy, I.; Mbefo, M.K.; Jovicic, A.; Kiely, A.; Holton, J.L.; Lee, S.J.; Gitler, A.D.; Eliezer, D.; et al. The novel Parkinson’s disease linked mutation G51D attenuates in vitro aggregation and membrane binding of alpha-synuclein, and enhances its secretion and nuclear localization in cells. Hum. Mol. Genet. 2014, 23, 4491–4509. [Google Scholar] [CrossRef]
  69. Ghosh, D.; Sahay, S.; Ranjan, P.; Salot, S.; Mohite, G.M.; Singh, P.K.; Dwivedi, S.; Carvalho, E.; Banerjee, R.; Kumar, A.; et al. The newly discovered Parkinson’s disease associated Finnish mutation (A53E) attenuates alpha-synuclein aggregation and membrane binding. Biochemistry 2014, 53, 6419–6421. [Google Scholar] [CrossRef] [PubMed]
  70. Jensen, P.H.; Nielsen, M.S.; Jakes, R.; Dotti, C.G.; Goedert, M. Binding of alpha-synuclein to brain vesicles is abolished by familial Parkinson’s disease mutation. J. Biol. Chem. 1998, 273, 26292–26294. [Google Scholar] [CrossRef] [Green Version]
  71. Jo, E.; Fuller, N.; Rand, R.P.; St George-Hyslop, P.; Fraser, P.E. Defective membrane interactions of familial Parkinson’s disease mutant A30P alpha-synuclein. J. Mol. Biol. 2002, 315, 799–807. [Google Scholar] [CrossRef]
  72. Tsigelny, I.F.; Sharikov, Y.; Kouznetsova, V.L.; Greenberg, J.P.; Wrasidlo, W.; Overk, C.; Gonzalez, T.; Trejo, M.; Spencer, B.; Kosberg, K.; et al. Molecular determinants of alpha-synuclein mutants’ oligomerization and membrane interactions. ACS Chem. Neurosci. 2015, 6, 403–416. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Eliezer, D.; Kutluay, E.; Bussell, R., Jr.; Browne, G. Conformational properties of alpha-synuclein in its free and lipid-associated states. J. Mol. Biol. 2001, 307, 1061–1073. [Google Scholar] [CrossRef] [PubMed]
  74. Weinreb, P.H.; Zhen, W.; Poon, A.W.; Conway, K.A.; Lansbury, P.T., Jr. NACP, a protein implicated in Alzheimer’s disease and learning, is natively unfolded. Biochemistry 1996, 35, 13709–13715. [Google Scholar] [CrossRef] [PubMed]
  75. Killinger, B.A.; Melki, R.; Brundin, P.; Kordower, J.H. Endogenous alpha-synuclein monomers, oligomers and resulting pathology: Let’s talk about the lipids in the room. NPJ Parkinsons Dis. 2019, 5, 23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Fauvet, B.; Mbefo, M.K.; Fares, M.B.; Desobry, C.; Michael, S.; Ardah, M.T.; Tsika, E.; Coune, P.; Prudent, M.; Lion, N.; et al. alpha-Synuclein in central nervous system and from erythrocytes, mammalian cells, and Escherichia coli exists predominantly as disordered monomer. J. Biol. Chem. 2012, 287, 15345–15364. [Google Scholar] [CrossRef] [Green Version]
  77. Bartels, T.; Choi, J.G.; Selkoe, D.J. alpha-Synuclein occurs physiologically as a helically folded tetramer that resists aggregation. Nature 2011, 477, 107–110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Wang, W.; Perovic, I.; Chittuluru, J.; Kaganovich, A.; Nguyen, L.T.; Liao, J.; Auclair, J.R.; Johnson, D.; Landeru, A.; Simorellis, A.K.; et al. A soluble alpha-synuclein construct forms a dynamic tetramer. Proc. Natl. Acad. Sci. USA 2011, 108, 17797–17802. [Google Scholar] [CrossRef] [Green Version]
  79. Dettmer, U.; Newman, A.J.; Luth, E.S.; Bartels, T.; Selkoe, D. In vivo cross-linking reveals principally oligomeric forms of alpha-synuclein and beta-synuclein in neurons and non-neural cells. J. Biol. Chem. 2013, 288, 6371–6385. [Google Scholar] [CrossRef] [Green Version]
  80. Davidson, W.S.; Jonas, A.; Clayton, D.F.; George, J.M. Stabilization of alpha-synuclein secondary structure upon binding to synthetic membranes. J. Biol. Chem. 1998, 273, 9443–9449. [Google Scholar] [CrossRef] [Green Version]
  81. Jo, E.; McLaurin, J.; Yip, C.M.; St George-Hyslop, P.; Fraser, P.E. Alpha-Synuclein membrane interactions and lipid specificity. J. Biol. Chem. 2000, 275, 34328–34334. [Google Scholar] [CrossRef] [Green Version]
  82. Trexler, A.J.; Rhoades, E. Alpha-synuclein binds large unilamellar vesicles as an extended helix. Biochemistry 2009, 48, 2304–2306. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Jao, C.C.; Der-Sarkissian, A.; Chen, J.; Langen, R. Structure of membrane-bound alpha-synuclein studied by site-directed spin labeling. Proc. Natl. Acad. Sci. USA 2004, 101, 8331–8336. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Chandra, S.; Chen, X.; Rizo, J.; Jahn, R.; Sudhof, T.C. A broken alpha -helix in folded alpha -Synuclein. J. Biol. Chem. 2003, 278, 15313–15318. [Google Scholar] [CrossRef] [Green Version]
  85. Drescher, M.; Veldhuis, G.; van Rooijen, B.D.; Milikisyants, S.; Subramaniam, V.; Huber, M. Antiparallel arrangement of the helices of vesicle-bound alpha-synuclein. J. Am. Chem. Soc. 2008, 130, 7796–7797. [Google Scholar] [CrossRef]
  86. Sharon, R.; Goldberg, M.S.; Bar-Josef, I.; Betensky, R.A.; Shen, J.; Selkoe, D.J. alpha-Synuclein occurs in lipid-rich high molecular weight complexes, binds fatty acids, and shows homology to the fatty acid-binding proteins. Proc. Natl. Acad. Sci. USA 2001, 98, 9110–9115. [Google Scholar] [CrossRef] [Green Version]
  87. Lucke, C.; Gantz, D.L.; Klimtchuk, E.; Hamilton, J.A. Interactions between fatty acids and alpha-synuclein. J. Lipid Res. 2006, 47, 1714–1724. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Madine, J.; Doig, A.J.; Middleton, D.A. A study of the regional effects of alpha-synuclein on the organization and stability of phospholipid bilayers. Biochemistry 2006, 45, 5783–5792. [Google Scholar] [CrossRef]
  89. Adamczyk, A.; Kacprzak, M.; Kazmierczak, A. Alpha-synuclein decreases arachidonic acid incorporation into rat striatal synaptoneurosomes. Folia Neuropathol. 2007, 45, 230–235. [Google Scholar] [PubMed]
  90. Castagnet, P.I.; Golovko, M.Y.; Barcelo-Coblijn, G.C.; Nussbaum, R.L.; Murphy, E.J. Fatty acid incorporation is decreased in astrocytes cultured from alpha-synuclein gene-ablated mice. J. Neurochem. 2005, 94, 839–849. [Google Scholar] [CrossRef]
  91. Barcelo-Coblijn, G.; Golovko, M.Y.; Weinhofer, I.; Berger, J.; Murphy, E.J. Brain neutral lipids mass is increased in alpha-synuclein gene-ablated mice. J. Neurochem. 2007, 101, 132–141. [Google Scholar] [CrossRef]
  92. Alza, N.P.; Conde, M.A.; Scodelaro-Bilbao, P.G.; Salvador, G.A. Neutral lipids as early biomarkers of cellular fate: The case of alpha-synuclein overexpression. Cell Death Dis. 2021, 12, 52. [Google Scholar] [CrossRef]
  93. Ruf, V.C.; Nubling, G.S.; Willikens, S.; Shi, S.; Schmidt, F.; Levin, J.; Botzel, K.; Kamp, F.; Giese, A. Different effects of alpha-synuclein mutants on lipid binding and aggregation detected by single molecule fluorescence spectroscopy and ThT fluorescence-based measurements. ACS Chem. Neurosci. 2019, 10, 1649–1659. [Google Scholar] [CrossRef]
  94. Conde, M.A.; Alza, N.P.; Iglesias Gonzalez, P.A.; Scodelaro Bilbao, P.G.; Sanchez Campos, S.; Uranga, R.M.; Salvador, G.A. Phospholipase D1 downregulation by alpha-synuclein: Implications for neurodegeneration in Parkinson’s disease. Biochim. Biophys. Acta Mol. Cell Biol. Lipids 2018, 1863, 639–650. [Google Scholar] [CrossRef]
  95. Ahn, B.H.; Rhim, H.; Kim, S.Y.; Sung, Y.M.; Lee, M.Y.; Choi, J.Y.; Wolozin, B.; Chang, J.S.; Lee, Y.H.; Kwon, T.K.; et al. alpha-Synuclein interacts with phospholipase D isozymes and inhibits pervanadate-induced phospholipase D activation in human embryonic kidney-293 cells. J. Biol. Chem. 2002, 277, 12334–12342. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Jenco, J.M.; Rawlingson, A.; Daniels, B.; Morris, A.J. Regulation of phospholipase D2: Selective inhibition of mammalian phospholipase D isoenzymes by alpha- and beta-synucleins. Biochemistry 1998, 37, 4901–4909. [Google Scholar] [CrossRef] [PubMed]
  97. Payton, J.E.; Perrin, R.J.; Woods, W.S.; George, J.M. Structural determinants of PLD2 inhibition by alpha-synuclein. J. Mol. Biol. 2004, 337, 1001–1009. [Google Scholar] [CrossRef]
  98. Rappley, I.; Gitler, A.D.; Selvy, P.E.; LaVoie, M.J.; Levy, B.D.; Brown, H.A.; Lindquist, S.; Selkoe, D.J. Evidence that alpha-synuclein does not inhibit phospholipase D. Biochemistry 2009, 48, 1077–1083. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Ulmer, T.S.; Bax, A.; Cole, N.B.; Nussbaum, R.L. Structure and dynamics of micelle-bound human alpha-synuclein. J. Biol. Chem. 2005, 280, 9595–9603. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Sode, K.; Ochiai, S.; Kobayashi, N.; Usuzaka, E. Effect of reparation of repeat sequences in the human alpha-synuclein on fibrillation ability. Int. J. Biol. Sci. 2006, 3, 1–7. [Google Scholar] [CrossRef] [Green Version]
  101. Giasson, B.I.; Murray, I.V.; Trojanowski, J.Q.; Lee, V.M. A hydrophobic stretch of 12 amino acid residues in the middle of alpha-synuclein is essential for filament assembly. J. Biol. Chem. 2001, 276, 2380–2386. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Rodriguez, J.A.; Ivanova, M.I.; Sawaya, M.R.; Cascio, D.; Reyes, F.E.; Shi, D.; Sangwan, S.; Guenther, E.L.; Johnson, L.M.; Zhang, M.; et al. Structure of the toxic core of alpha-synuclein from invisible crystals. Nature 2015, 525, 486–490. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Longhena, F.; Faustini, G.; Missale, C.; Pizzi, M.; Spano, P.; Bellucci, A. The Contribution of alpha-Synuclein Spreading to Parkinson’s Disease Synaptopathy. Neural. Plast. 2017, 2017, 5012129. [Google Scholar] [CrossRef] [Green Version]
  104. Volpicelli-Daley, L.A.; Luk, K.C.; Patel, T.P.; Tanik, S.A.; Riddle, D.M.; Stieber, A.; Meaney, D.F.; Trojanowski, J.Q.; Lee, V.M. Exogenous alpha-synuclein fibrils induce Lewy body pathology leading to synaptic dysfunction and neuron death. Neuron 2011, 72, 57–71. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Appel-Cresswell, S.; Vilarino-Guell, C.; Encarnacion, M.; Sherman, H.; Yu, I.; Shah, B.; Weir, D.; Thompson, C.; Szu-Tu, C.; Trinh, J.; et al. Alpha-synuclein p.H50Q, a novel pathogenic mutation for Parkinson’s disease. Mov. Disord. 2013, 28, 811–813. [Google Scholar] [CrossRef]
  106. Kiely, A.P.; Asi, Y.T.; Kara, E.; Limousin, P.; Ling, H.; Lewis, P.; Proukakis, C.; Quinn, N.; Lees, A.J.; Hardy, J.; et al. alpha-Synucleinopathy associated with G51D SNCA mutation: A link between Parkinson’s disease and multiple system atrophy? Acta Neuropathol. 2013, 125, 753–769. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Hoffman-Zacharska, D.; Koziorowski, D.; Ross, O.A.; Milewski, M.; Poznanski, J.A.; Jurek, M.; Wszolek, Z.K.; Soto-Ortolaza, A.; Awek, J.A.S.; Janik, P.; et al. Novel A18T and pA29S substitutions in alpha-synuclein may be associated with sporadic Parkinson’s disease. Parkinsonism Relat. Disord. 2013, 19, 1057–1060. [Google Scholar] [CrossRef] [Green Version]
  108. Conway, K.A.; Lee, S.J.; Rochet, J.C.; Ding, T.T.; Williamson, R.E.; Lansbury, P.T., Jr. Acceleration of oligomerization, not fibrillization, is a shared property of both alpha-synuclein mutations linked to early-onset Parkinson’s disease: Implications for pathogenesis and therapy. Proc. Natl. Acad. Sci. USA 2000, 97, 571–576. [Google Scholar] [CrossRef] [Green Version]
  109. Conway, K.A.; Harper, J.D.; Lansbury, P.T. Accelerated in vitro fibril formation by a mutant alpha-synuclein linked to early-onset Parkinson disease. Nat. Med. 1998, 4, 1318–1320. [Google Scholar] [CrossRef] [PubMed]
  110. Bussell, R., Jr.; Eliezer, D. Residual structure and dynamics in Parkinson’s disease-associated mutants of alpha-synuclein. J. Biol. Chem. 2001, 276, 45996–46003. [Google Scholar] [CrossRef] [Green Version]
  111. Lazaro, D.F.; Rodrigues, E.F.; Langohr, R.; Shahpasandzadeh, H.; Ribeiro, T.; Guerreiro, P.; Gerhardt, E.; Krohnert, K.; Klucken, J.; Pereira, M.D.; et al. Systematic comparison of the effects of alpha-synuclein mutations on its oligomerization and aggregation. PLoS Genet. 2014, 10, e1004741. [Google Scholar] [CrossRef]
  112. Fredenburg, R.A.; Rospigliosi, C.; Meray, R.K.; Kessler, J.C.; Lashuel, H.A.; Eliezer, D.; Lansbury, P.T., Jr. The impact of the E46K mutation on the properties of alpha-synuclein in its monomeric and oligomeric states. Biochemistry 2007, 46, 7107–7118. [Google Scholar] [CrossRef] [Green Version]
  113. Rospigliosi, C.C.; McClendon, S.; Schmid, A.W.; Ramlall, T.F.; Barre, P.; Lashuel, H.A.; Eliezer, D. E46K Parkinson’s-linked mutation enhances C-terminal-to-N-terminal contacts in alpha-synuclein. J. Mol. Biol. 2009, 388, 1022–1032. [Google Scholar] [CrossRef] [Green Version]
  114. Khalaf, O.; Fauvet, B.; Oueslati, A.; Dikiy, I.; Mahul-Mellier, A.L.; Ruggeri, F.S.; Mbefo, M.K.; Vercruysse, F.; Dietler, G.; Lee, S.J.; et al. The H50Q mutation enhances alpha-synuclein aggregation, secretion, and toxicity. J. Biol. Chem. 2014, 289, 21856–21876. [Google Scholar] [CrossRef] [Green Version]
  115. Sanjeev, A.; Mattaparthi, V.S.K. Computational investigation on the effects of H50Q and G51D mutations on the alpha-Synuclein aggregation propensity. J. Biomol. Struct. Dyn. 2018, 36, 2224–2236. [Google Scholar] [CrossRef]
  116. Lesage, S.; Anheim, M.; Letournel, F.; Bousset, L.; Honore, A.; Rozas, N.; Pieri, L.; Madiona, K.; Durr, A.; Melki, R.; et al. G51D alpha-synuclein mutation causes a novel parkinsonian-pyramidal syndrome. Ann. Neurol. 2013, 73, 459–471. [Google Scholar] [CrossRef] [PubMed]
  117. Rutherford, N.J.; Moore, B.D.; Golde, T.E.; Giasson, B.I. Divergent effects of the H50Q and G51D SNCA mutations on the aggregation of alpha-synuclein. J. Neurochem. 2014, 131, 859–867. [Google Scholar] [CrossRef]
  118. Rutherford, N.J.; Giasson, B.I. The A53E alpha-synuclein pathological mutation demonstrates reduced aggregation propensity in vitro and in cell culture. Neurosci. Lett. 2015, 597, 43–48. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Pasanen, P.; Myllykangas, L.; Siitonen, M.; Raunio, A.; Kaakkola, S.; Lyytinen, J.; Tienari, P.J.; Poyhonen, M.; Paetau, A. Novel alpha-synuclein mutation A53E associated with atypical multiple system atrophy and Parkinson’s disease-type pathology. Neurobiol. Aging 2014, 35, 2180.e1. [Google Scholar] [CrossRef] [PubMed]
  120. Ozawa, T.; Takano, H.; Onodera, O.; Kobayashi, H.; Ikeuchi, T.; Koide, R.; Okuizumi, K.; Shimohata, T.; Wakabayashi, K.; Takahashi, H.; et al. No mutation in the entire coding region of the alpha-synuclein gene in pathologically confirmed cases of multiple system atrophy. Neurosci. Lett. 1999, 270, 110–112. [Google Scholar] [CrossRef]
  121. Narhi, L.; Wood, S.J.; Steavenson, S.; Jiang, Y.; Wu, G.M.; Anafi, D.; Kaufman, S.A.; Martin, F.; Sitney, K.; Denis, P.; et al. Both familial Parkinson’s disease mutations accelerate alpha-synuclein aggregation. J. Biol. Chem. 1999, 274, 9843–9846. [Google Scholar] [CrossRef] [Green Version]
  122. Pandey, N.; Schmidt, R.E.; Galvin, J.E. The alpha-synuclein mutation E46K promotes aggregation in cultured cells. Exp. Neurol. 2006, 197, 515–520. [Google Scholar] [CrossRef] [PubMed]
  123. Smith, W.W.; Margolis, R.L.; Li, X.; Troncoso, J.C.; Lee, M.K.; Dawson, V.L.; Dawson, T.M.; Iwatsubo, T.; Ross, C.A. Alpha-synuclein phosphorylation enhances eosinophilic cytoplasmic inclusion formation in SH-SY5Y cells. J. Neurosci. 2005, 25, 5544–5552. [Google Scholar] [CrossRef] [PubMed]
  124. Sugeno, N.; Takeda, A.; Hasegawa, T.; Kobayashi, M.; Kikuchi, A.; Mori, F.; Wakabayashi, K.; Itoyama, Y. Serine 129 phosphorylation of alpha-synuclein induces unfolded protein response-mediated cell death. J. Biol. Chem. 2008, 283, 23179–23188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Ma, M.R.; Hu, Z.W.; Zhao, Y.F.; Chen, Y.X.; Li, Y.M. Phosphorylation induces distinct alpha-synuclein strain formation. Sci. Rep. 2016, 6, 37130. [Google Scholar] [CrossRef]
  126. Sato, H.; Arawaka, S.; Hara, S.; Fukushima, S.; Koga, K.; Koyama, S.; Kato, T. Authentically phosphorylated alpha-synuclein at Ser129 accelerates neurodegeneration in a rat model of familial Parkinson’s disease. J. Neurosci. 2011, 31, 16884–16894. [Google Scholar] [CrossRef]
  127. Chen, L.; Periquet, M.; Wang, X.; Negro, A.; McLean, P.J.; Hyman, B.T.; Feany, M.B. Tyrosine and serine phosphorylation of alpha-synuclein have opposing effects on neurotoxicity and soluble oligomer formation. J. Clin. Investig. 2009, 119, 3257–3265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Azeredo da Silveira, S.; Schneider, B.L.; Cifuentes-Diaz, C.; Sage, D.; Abbas-Terki, T.; Iwatsubo, T.; Unser, M.; Aebischer, P. Phosphorylation does not prompt, nor prevent, the formation of alpha-synuclein toxic species in a rat model of Parkinson’s disease. Hum. Mol. Genet. 2009, 18, 872–887. [Google Scholar] [CrossRef] [Green Version]
  129. Mbefo, M.K.; Paleologou, K.E.; Boucharaba, A.; Oueslati, A.; Schell, H.; Fournier, M.; Olschewski, D.; Yin, G.; Zweckstetter, M.; Masliah, E.; et al. Phosphorylation of synucleins by members of the Polo-like kinase family. J. Biol. Chem. 2010, 285, 2807–2822. [Google Scholar] [CrossRef] [Green Version]
  130. Chen, L.; Feany, M.B. Alpha-synuclein phosphorylation controls neurotoxicity and inclusion formation in a Drosophila model of Parkinson disease. Nat. Neurosci. 2005, 8, 657–663. [Google Scholar] [CrossRef]
  131. Paleologou, K.E.; Schmid, A.W.; Rospigliosi, C.C.; Kim, H.Y.; Lamberto, G.R.; Fredenburg, R.A.; Lansbury, P.T., Jr.; Fernandez, C.O.; Eliezer, D.; Zweckstetter, M.; et al. Phosphorylation at Ser-129 but not the phosphomimics S129E/D inhibits the fibrillation of alpha-synuclein. J. Biol. Chem. 2008, 283, 16895–16905. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Wu, B.; Liu, Q.; Duan, C.; Li, Y.; Yu, S.; Chan, P.; Ueda, K.; Yang, H. Phosphorylation of alpha-synuclein upregulates tyrosine hydroxylase activity in MN9D cells. Acta Histochem. 2011, 113, 32–35. [Google Scholar] [CrossRef] [PubMed]
  133. Kuwahara, T.; Tonegawa, R.; Ito, G.; Mitani, S.; Iwatsubo, T. Phosphorylation of alpha-synuclein protein at Ser-129 reduces neuronal dysfunction by lowering its membrane binding property in Caenorhabditis elegans. J. Biol. Chem. 2012, 287, 7098–7109. [Google Scholar] [CrossRef] [Green Version]
  134. Anderson, J.P.; Walker, D.E.; Goldstein, J.M.; de Laat, R.; Banducci, K.; Caccavello, R.J.; Barbour, R.; Huang, J.; Kling, K.; Lee, M.; et al. Phosphorylation of Ser-129 is the dominant pathological modification of alpha-synuclein in familial and sporadic Lewy body disease. J. Biol. Chem. 2006, 281, 29739–29752. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Fujiwara, H.; Hasegawa, M.; Dohmae, N.; Kawashima, A.; Masliah, E.; Goldberg, M.S.; Shen, J.; Takio, K.; Iwatsubo, T. alpha-Synuclein is phosphorylated in synucleinopathy lesions. Nat. Cell Biol. 2002, 4, 160–164. [Google Scholar] [CrossRef] [PubMed]
  136. Sato, H.; Kato, T.; Arawaka, S. The role of Ser129 phosphorylation of alpha-synuclein in neurodegeneration of Parkinson’s disease: A review of in vivo models. Rev. Neurosci. 2013, 24, 115–123. [Google Scholar] [CrossRef]
  137. Walker, D.G.; Lue, L.F.; Adler, C.H.; Shill, H.A.; Caviness, J.N.; Sabbagh, M.N.; Akiyama, H.; Serrano, G.E.; Sue, L.I.; Beach, T.G. Changes in properties of serine 129 phosphorylated alpha-synuclein with progression of Lewy-type histopathology in human brains. Exp. Neurol. 2013, 240, 190–204. [Google Scholar] [CrossRef] [Green Version]
  138. Paleologou, K.E.; Oueslati, A.; Shakked, G.; Rospigliosi, C.C.; Kim, H.Y.; Lamberto, G.R.; Fernandez, C.O.; Schmid, A.; Chegini, F.; Gai, W.P.; et al. Phosphorylation at S87 is enhanced in synucleinopathies, inhibits alpha-synuclein oligomerization, and influences synuclein-membrane interactions. J. Neurosci. 2010, 30, 3184–3198. [Google Scholar] [CrossRef]
  139. Xu, Y.; Deng, Y.; Qing, H. The phosphorylation of alpha-synuclein: Development and implication for the mechanism and therapy of the Parkinson’s disease. J. Neurochem. 2015, 135, 4–18. [Google Scholar] [CrossRef] [Green Version]
  140. Lu, Y.; Prudent, M.; Fauvet, B.; Lashuel, H.A.; Girault, H.H. Phosphorylation of alpha-Synuclein at Y125 and S129 alters its metal binding properties: Implications for understanding the role of alpha-Synuclein in the pathogenesis of Parkinson’s Disease and related disorders. ACS Chem. Neurosci. 2011, 2, 667–675. [Google Scholar] [CrossRef] [Green Version]
  141. Fayyad, M.; Erskine, D.; Majbour, N.K.; Vaikath, N.N.; Ghanem, S.S.; Sudhakaran, I.P.; Abdesselem, H.; Lamprokostopoulou, A.; Vekrellis, K.; Morris, C.M.; et al. Investigating the presence of doubly phosphorylated alpha-synuclein at tyrosine 125 and serine 129 in idiopathic Lewy body diseases. Brain Pathol. 2020, 30, 831–843. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Giasson, B.I.; Duda, J.E.; Murray, I.V.; Chen, Q.; Souza, J.M.; Hurtig, H.I.; Ischiropoulos, H.; Trojanowski, J.Q.; Lee, V.M. Oxidative damage linked to neurodegeneration by selective alpha-synuclein nitration in synucleinopathy lesions. Science 2000, 290, 985–989. [Google Scholar] [CrossRef] [PubMed]
  143. Danielson, S.R.; Held, J.M.; Schilling, B.; Oo, M.; Gibson, B.W.; Andersen, J.K. Preferentially increased nitration of alpha-synuclein at tyrosine-39 in a cellular oxidative model of Parkinson’s disease. Anal. Chem. 2009, 81, 7823–7828. [Google Scholar] [CrossRef] [Green Version]
  144. McCormack, A.L.; Mak, S.K.; Di Monte, D.A. Increased alpha-synuclein phosphorylation and nitration in the aging primate substantia nigra. Cell Death Dis. 2012, 3, e315. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Yu, Z.; Xu, X.; Xiang, Z.; Zhou, J.; Zhang, Z.; Hu, C.; He, C. Nitrated alpha-synuclein induces the loss of dopaminergic neurons in the substantia nigra of rats. PLoS ONE 2010, 5, e9956. [Google Scholar] [CrossRef]
  146. Souza, J.M.; Peluffo, G.; Radi, R. Protein tyrosine nitration—Functional alteration or just a biomarker? Free Radic. Biol. Med. 2008, 45, 357–366. [Google Scholar] [CrossRef]
  147. Hodara, R.; Norris, E.H.; Giasson, B.I.; Mishizen-Eberz, A.J.; Lynch, D.R.; Lee, V.M.; Ischiropoulos, H. Functional consequences of alpha-synuclein tyrosine nitration: Diminished binding to lipid vesicles and increased fibril formation. J. Biol. Chem. 2004, 279, 47746–47753. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Takahashi, T.; Yamashita, H.; Nakamura, T.; Nagano, Y.; Nakamura, S. Tyrosine 125 of alpha-synuclein plays a critical role for dimerization following nitrative stress. Brain Res. 2002, 938, 73–80. [Google Scholar] [CrossRef]
  149. Souza, J.M.; Giasson, B.I.; Chen, Q.; Lee, V.M.; Ischiropoulos, H. Dityrosine cross-linking promotes formation of stable alpha -synuclein polymers. Implication of nitrative and oxidative stress in the pathogenesis of neurodegenerative synucleinopathies. J. Biol. Chem. 2000, 275, 18344–18349. [Google Scholar] [CrossRef] [Green Version]
  150. Fernandez, E.; Garcia-Moreno, J.M.; Martin de Pablos, A.; Chacon, J. May the evaluation of nitrosative stress through selective increase of 3-nitrotyrosine proteins other than nitroalbumin and dominant tyrosine-125/136 nitrosylation of serum alpha-synuclein serve for diagnosis of sporadic Parkinson’s disease? Antioxid. Redox. Signal. 2013, 19, 912–918. [Google Scholar] [CrossRef] [Green Version]
  151. Grice, G.L.; Nathan, J.A. The recognition of ubiquitinated proteins by the proteasome. Cell Mol. Life Sci. 2016, 73, 3497–3506. [Google Scholar] [CrossRef] [Green Version]
  152. Pickart, C.M.; Eddins, M.J. Ubiquitin: Structures, functions, mechanisms. Biochim. Biophys. Acta 2004, 1695, 55–72. [Google Scholar] [CrossRef] [Green Version]
  153. Tofaris, G.K.; Razzaq, A.; Ghetti, B.; Lilley, K.S.; Spillantini, M.G. Ubiquitination of alpha-synuclein in Lewy bodies is a pathological event not associated with impairment of proteasome function. J. Biol. Chem. 2003, 278, 44405–44411. [Google Scholar] [CrossRef] [Green Version]
  154. Hasegawa, M.; Fujiwara, H.; Nonaka, T.; Wakabayashi, K.; Takahashi, H.; Lee, V.M.; Trojanowski, J.Q.; Mann, D.; Iwatsubo, T. Phosphorylated alpha-synuclein is ubiquitinated in alpha-synucleinopathy lesions. J. Biol. Chem. 2002, 277, 49071–49076. [Google Scholar] [CrossRef] [Green Version]
  155. Liani, E.; Eyal, A.; Avraham, E.; Shemer, R.; Szargel, R.; Berg, D.; Bornemann, A.; Riess, O.; Ross, C.A.; Rott, R.; et al. Ubiquitylation of synphilin-1 and alpha-synuclein by SIAH and its presence in cellular inclusions and Lewy bodies imply a role in Parkinson’s disease. Proc. Natl. Acad. Sci. USA 2004, 101, 5500–5505. [Google Scholar] [CrossRef] [Green Version]
  156. Shin, Y.; Klucken, J.; Patterson, C.; Hyman, B.T.; McLean, P.J. The co-chaperone carboxyl terminus of Hsp70-interacting protein (CHIP) mediates alpha-synuclein degradation decisions between proteasomal and lysosomal pathways. J. Biol. Chem. 2005, 280, 23727–23734. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Tofaris, G.K.; Kim, H.T.; Hourez, R.; Jung, J.W.; Kim, K.P.; Goldberg, A.L. Ubiquitin ligase Nedd4 promotes alpha-synuclein degradation by the endosomal-lysosomal pathway. Proc. Natl. Acad. Sci. USA 2011, 108, 17004–17009. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Lee, J.T.; Wheeler, T.C.; Li, L.; Chin, L.S. Ubiquitination of alpha-synuclein by Siah-1 promotes alpha-synuclein aggregation and apoptotic cell death. Hum. Mol. Genet. 2008, 17, 906–917. [Google Scholar] [CrossRef] [PubMed]
  159. Rott, R.; Szargel, R.; Haskin, J.; Shani, V.; Shainskaya, A.; Manov, I.; Liani, E.; Avraham, E.; Engelender, S. Monoubiquitylation of alpha-synuclein by seven in absentia homolog (SIAH) promotes its aggregation in dopaminergic cells. J. Biol. Chem. 2008, 283, 3316–3328. [Google Scholar] [CrossRef] [Green Version]
  160. Kalia, L.V.; Kalia, S.K.; Chau, H.; Lozano, A.M.; Hyman, B.T.; McLean, P.J. Ubiquitinylation of alpha-synuclein by carboxyl terminus Hsp70-interacting protein (CHIP) is regulated by Bcl-2-associated athanogene 5 (BAG5). PLoS ONE 2011, 6, e14695. [Google Scholar] [CrossRef] [PubMed]
  161. Meier, F.; Abeywardana, T.; Dhall, A.; Marotta, N.P.; Varkey, J.; Langen, R.; Chatterjee, C.; Pratt, M.R. Semisynthetic, site-specific ubiquitin modification of alpha-synuclein reveals differential effects on aggregation. J. Am. Chem. Soc. 2012, 134, 5468–5471. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Weetman, J.; Wong, M.B.; Sharry, S.; Rcom-H’cheo-Gauthier, A.; Gai, W.P.; Meedeniya, A.; Pountney, D.L. Increased SUMO-1 expression in the unilateral rotenone-lesioned mouse model of Parkinson’s disease. Neurosci. Lett. 2013, 544, 119–124. [Google Scholar] [CrossRef] [PubMed]
  163. Wong, M.B.; Goodwin, J.; Norazit, A.; Meedeniya, A.C.; Richter-Landsberg, C.; Gai, W.P.; Pountney, D.L. SUMO-1 is associated with a subset of lysosomes in glial protein aggregate diseases. Neurotox. Res. 2013, 23, 1–21. [Google Scholar] [CrossRef]
  164. Pountney, D.L.; Chegini, F.; Shen, X.; Blumbergs, P.C.; Gai, W.P. SUMO-1 marks subdomains within glial cytoplasmic inclusions of multiple system atrophy. Neurosci. Lett. 2005, 381, 74–79. [Google Scholar] [CrossRef]
  165. Rott, R.; Szargel, R.; Shani, V.; Hamza, H.; Savyon, M.; Abd Elghani, F.; Bandopadhyay, R.; Engelender, S. SUMOylation and ubiquitination reciprocally regulate alpha-synuclein degradation and pathological aggregation. Proc. Natl. Acad. Sci. USA 2017, 114, 13176–13181. [Google Scholar] [CrossRef] [Green Version]
  166. Abeywardana, T.; Pratt, M.R. Extent of inhibition of alpha-synuclein aggregation in vitro by SUMOylation is conjugation site- and SUMO isoform-selective. Biochemistry 2015, 54, 959–961. [Google Scholar] [CrossRef] [PubMed]
  167. Krumova, P.; Meulmeester, E.; Garrido, M.; Tirard, M.; Hsiao, H.H.; Bossis, G.; Urlaub, H.; Zweckstetter, M.; Kugler, S.; Melchior, F.; et al. Sumoylation inhibits alpha-synuclein aggregation and toxicity. J. Cell Biol. 2011, 194, 49–60. [Google Scholar] [CrossRef] [Green Version]
  168. Vinueza-Gavilanes, R.; Inigo-Marco, I.; Larrea, L.; Lasa, M.; Carte, B.; Santamaria, E.; Fernandez-Irigoyen, J.; Bugallo, R.; Aragon, T.; Aldabe, R.; et al. N-terminal acetylation mutants affect alpha-synuclein stability, protein levels and neuronal toxicity. Neurobiol. Dis. 2020, 137, 104781. [Google Scholar] [CrossRef] [PubMed]
  169. Iyer, A.; Roeters, S.J.; Schilderink, N.; Hommersom, B.; Heeren, R.M.; Woutersen, S.; Claessens, M.M.; Subramaniam, V. The impact of N-terminal acetylation of alpha-synuclein on phospholipid membrane binding and fibril structure. J. Biol. Chem. 2016, 291, 21110–21122. [Google Scholar] [CrossRef] [Green Version]
  170. Kang, L.; Moriarty, G.M.; Woods, L.A.; Ashcroft, A.E.; Radford, S.E.; Baum, J. N-terminal acetylation of alpha-synuclein induces increased transient helical propensity and decreased aggregation rates in the intrinsically disordered monomer. Protein Sci. 2012, 21, 911–917. [Google Scholar] [CrossRef] [Green Version]
  171. Runfola, M.; De Simone, A.; Vendruscolo, M.; Dobson, C.M.; Fusco, G. The N-terminal acetylation of alpha-synuclein changes the affinity for lipid membranes but not the structural properties of the bound state. Sci. Rep. 2020, 10, 204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Bu, B.; Tong, X.; Li, D.; Hu, Y.; He, W.; Zhao, C.; Hu, R.; Li, X.; Shao, Y.; Liu, C.; et al. N-terminal acetylation preserves alpha-synuclein from oligomerization by blocking intermolecular hydrogen bonds. ACS Chem. Neurosci. 2017, 8, 2145–2151. [Google Scholar] [CrossRef] [PubMed]
  173. Bartels, T.; Kim, N.C.; Luth, E.S.; Selkoe, D.J. N-alpha-acetylation of alpha-synuclein increases its helical folding propensity, GM1 binding specificity and resistance to aggregation. PLoS ONE 2014, 9, e103727. [Google Scholar] [CrossRef] [Green Version]
  174. Wang, S.; Yang, F.; Petyuk, V.A.; Shukla, A.K.; Monroe, M.E.; Gritsenko, M.A.; Rodland, K.D.; Smith, R.D.; Qian, W.J.; Gong, C.X.; et al. Quantitative proteomics identifies altered O-GlcNAcylation of structural, synaptic and memory-associated proteins in Alzheimer’s disease. J. Pathol. 2017, 243, 78–88. [Google Scholar] [CrossRef] [Green Version]
  175. Alfaro, J.F.; Gong, C.X.; Monroe, M.E.; Aldrich, J.T.; Clauss, T.R.; Purvine, S.O.; Wang, Z.; Camp, D.G., 2nd; Shabanowitz, J.; Stanley, P.; et al. Tandem mass spectrometry identifies many mouse brain O-GlcNAcylated proteins including EGF domain-specific O-GlcNAc transferase targets. Proc. Natl. Acad. Sci. USA 2012, 109, 7280–7285. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Morris, M.; Knudsen, G.M.; Maeda, S.; Trinidad, J.C.; Ioanoviciu, A.; Burlingame, A.L.; Mucke, L. Tau post-translational modifications in wild-type and human amyloid precursor protein transgenic mice. Nat. Neurosci. 2015, 18, 1183–1189. [Google Scholar] [CrossRef] [PubMed]
  177. Wang, Z.; Park, K.; Comer, F.; Hsieh-Wilson, L.C.; Saudek, C.D.; Hart, G.W. Site-specific GlcNAcylation of human erythrocyte proteins: Potential biomarker(s) for diabetes. Diabetes 2009, 58, 309–317. [Google Scholar] [CrossRef] [Green Version]
  178. Wang, Z.; Udeshi, N.D.; O’Malley, M.; Shabanowitz, J.; Hunt, D.F.; Hart, G.W. Enrichment and site mapping of O-linked N-acetylglucosamine by a combination of chemical/enzymatic tagging, photochemical cleavage, and electron transfer dissociation mass spectrometry. Mol. Cell Proteom. 2010, 9, 153–160. [Google Scholar] [CrossRef] [Green Version]
  179. Zhang, J.; Lei, H.; Chen, Y.; Ma, Y.T.; Jiang, F.; Tan, J.; Zhang, Y.; Li, J.D. Enzymatic O-GlcNAcylation of alpha-synuclein reduces aggregation and increases SDS-resistant soluble oligomers. Neurosci. Lett. 2017, 655, 90–94. [Google Scholar] [CrossRef]
  180. Levine, P.M.; Galesic, A.; Balana, A.T.; Mahul-Mellier, A.L.; Navarro, M.X.; De Leon, C.A.; Lashuel, H.A.; Pratt, M.R. alpha-Synuclein O-GlcNAcylation alters aggregation and toxicity, revealing certain residues as potential inhibitors of Parkinson’s disease. Proc. Natl. Acad. Sci. USA 2019, 116, 1511–1519. [Google Scholar] [CrossRef] [Green Version]
  181. Lewis, Y.E.; Galesic, A.; Levine, P.M.; De Leon, C.A.; Lamiri, N.; Brennan, C.K.; Pratt, M.R. O-GlcNAcylation of alpha-synuclein at serine 87 reduces aggregation without affecting membrane binding. ACS Chem. Biol. 2017, 12, 1020–1027. [Google Scholar] [CrossRef] [PubMed]
  182. Marotta, N.P.; Lin, Y.H.; Lewis, Y.E.; Ambroso, M.R.; Zaro, B.W.; Roth, M.T.; Arnold, D.B.; Langen, R.; Pratt, M.R. O-GlcNAc modification blocks the aggregation and toxicity of the protein alpha-synuclein associated with Parkinson’s disease. Nat. Chem. 2015, 7, 913–920. [Google Scholar] [CrossRef] [PubMed]
  183. Li, W.; West, N.; Colla, E.; Pletnikova, O.; Troncoso, J.C.; Marsh, L.; Dawson, T.M.; Jakala, P.; Hartmann, T.; Price, D.L.; et al. Aggregation promoting C-terminal truncation of alpha-synuclein is a normal cellular process and is enhanced by the familial Parkinson’s disease-linked mutations. Proc. Natl. Acad. Sci. USA 2005, 102, 2162–2167. [Google Scholar] [CrossRef] [Green Version]
  184. Kellie, J.F.; Higgs, R.E.; Ryder, J.W.; Major, A.; Beach, T.G.; Adler, C.H.; Merchant, K.; Knierman, M.D. Quantitative measurement of intact alpha-synuclein proteoforms from post-mortem control and Parkinson’s disease brain tissue by intact protein mass spectrometry. Sci. Rep. 2014, 4, 5797. [Google Scholar] [CrossRef]
  185. Baba, M.; Nakajo, S.; Tu, P.H.; Tomita, T.; Nakaya, K.; Lee, V.M.; Trojanowski, J.Q.; Iwatsubo, T. Aggregation of alpha-synuclein in Lewy bodies of sporadic Parkinson’s disease and dementia with Lewy bodies. Am. J. Pathol. 1998, 152, 879–884. [Google Scholar]
  186. Ulusoy, A.; Febbraro, F.; Jensen, P.H.; Kirik, D.; Romero-Ramos, M. Co-expression of C-terminal truncated alpha-synuclein enhances full-length alpha-synuclein-induced pathology. Eur. J. Neurosci. 2010, 32, 409–422. [Google Scholar] [CrossRef] [PubMed]
  187. Hall, K.; Yang, S.; Sauchanka, O.; Spillantini, M.G.; Anichtchik, O. Behavioural deficits in transgenic mice expressing human truncated (1-120 amino acid) alpha-synuclein. Exp. Neurol. 2015, 264, 8–13. [Google Scholar] [CrossRef] [PubMed]
  188. Wang, W.; Nguyen, L.T.; Burlak, C.; Chegini, F.; Guo, F.; Chataway, T.; Ju, S.; Fisher, O.S.; Miller, D.W.; Datta, D.; et al. Caspase-1 causes truncation and aggregation of the Parkinson’s disease-associated protein alpha-synuclein. Proc. Natl. Acad. Sci. USA 2016, 113, 9587–9592. [Google Scholar] [CrossRef] [Green Version]
  189. Iyer, A.; Roeters, S.J.; Kogan, V.; Woutersen, S.; Claessens, M.; Subramaniam, V. C-terminal truncated alpha-synuclein fibrils contain strongly twisted beta-sheets. J. Am. Chem. Soc. 2017, 139, 15392–15400. [Google Scholar] [CrossRef]
  190. Ma, L.; Yang, C.; Zhang, X.; Li, Y.; Wang, S.; Zheng, L.; Huang, K. C-terminal truncation exacerbates the aggregation and cytotoxicity of alpha-Synuclein: A vicious cycle in Parkinson’s disease. Biochim. Biophys. Acta Mol. Basis Dis. 2018, 1864, 3714–3725. [Google Scholar] [CrossRef] [PubMed]
  191. Terada, M.; Suzuki, G.; Nonaka, T.; Kametani, F.; Tamaoka, A.; Hasegawa, M. The effect of truncation on prion-like properties of alpha-synuclein. J. Biol. Chem. 2018, 293, 13910–13920. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Tofaris, G.K.; Garcia Reitbock, P.; Humby, T.; Lambourne, S.L.; O’Connell, M.; Ghetti, B.; Gossage, H.; Emson, P.C.; Wilkinson, L.S.; Goedert, M.; et al. Pathological changes in dopaminergic nerve cells of the substantia nigra and olfactory bulb in mice transgenic for truncated human alpha-synuclein(1-120): Implications for Lewy body disorders. J. Neurosci. 2006, 26, 3942–3950. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Hoyer, W.; Cherny, D.; Subramaniam, V.; Jovin, T.M. Impact of the acidic C-terminal region comprising amino acids 109-140 on alpha-synuclein aggregation in vitro. Biochemistry 2004, 43, 16233–16242. [Google Scholar] [CrossRef] [PubMed]
  194. Murray, I.V.; Giasson, B.I.; Quinn, S.M.; Koppaka, V.; Axelsen, P.H.; Ischiropoulos, H.; Trojanowski, J.Q.; Lee, V.M. Role of alpha-synuclein carboxy-terminus on fibril formation in vitro. Biochemistry 2003, 42, 8530–8540. [Google Scholar] [CrossRef] [PubMed]
  195. Tsigelny, I.F.; Sharikov, Y.; Wrasidlo, W.; Gonzalez, T.; Desplats, P.A.; Crews, L.; Spencer, B.; Masliah, E. Role of alpha-synuclein penetration into the membrane in the mechanisms of oligomer pore formation. FEBS J. 2012, 279, 1000–1013. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Fantini, J.; Yahi, N. The driving force of alpha-synuclein insertion and amyloid channel formation in the plasma membrane of neural cells: Key role of ganglioside- and cholesterol-binding domains. Adv. Exp. Med. Biol. 2013, 991, 15–26. [Google Scholar] [CrossRef] [PubMed]
  197. Zakharov, S.D.; Hulleman, J.D.; Dutseva, E.A.; Antonenko, Y.N.; Rochet, J.C.; Cramer, W.A. Helical alpha-synuclein forms highly conductive ion channels. Biochemistry 2007, 46, 14369–14379. [Google Scholar] [CrossRef] [PubMed]
  198. Quist, A.; Doudevski, I.; Lin, H.; Azimova, R.; Ng, D.; Frangione, B.; Kagan, B.; Ghiso, J.; Lal, R. Amyloid ion channels: A common structural link for protein-misfolding disease. Proc. Natl. Acad. Sci. USA 2005, 102, 10427–10432. [Google Scholar] [CrossRef] [Green Version]
  199. Volles, M.J.; Lansbury, P.T., Jr. Vesicle permeabilization by protofibrillar alpha-synuclein is sensitive to Parkinson’s disease-linked mutations and occurs by a pore-like mechanism. Biochemistry 2002, 41, 4595–4602. [Google Scholar] [CrossRef]
  200. Stockl, M.T.; Zijlstra, N.; Subramaniam, V. alpha-Synuclein oligomers: An amyloid pore? Insights into mechanisms of alpha-synuclein oligomer-lipid interactions. Mol. Neurobiol. 2013, 47, 613–621. [Google Scholar] [CrossRef]
  201. Di Pasquale, E.; Fantini, J.; Chahinian, H.; Maresca, M.; Taieb, N.; Yahi, N. Altered ion channel formation by the Parkinson’s-disease-linked E46K mutant of alpha-synuclein is corrected by GM3 but not by GM1 gangliosides. J. Mol. Biol. 2010, 397, 202–218. [Google Scholar] [CrossRef] [PubMed]
  202. Furukawa, K.; Matsuzaki-Kobayashi, M.; Hasegawa, T.; Kikuchi, A.; Sugeno, N.; Itoyama, Y.; Wang, Y.; Yao, P.J.; Bushlin, I.; Takeda, A. Plasma membrane ion permeability induced by mutant alpha-synuclein contributes to the degeneration of neural cells. J. Neurochem. 2006, 97, 1071–1077. [Google Scholar] [CrossRef]
  203. Feng, L.R.; Federoff, H.J.; Vicini, S.; Maguire-Zeiss, K.A. Alpha-synuclein mediates alterations in membrane conductance: A potential role for alpha-synuclein oligomers in cell vulnerability. Eur. J. Neurosci. 2010, 32, 10–17. [Google Scholar] [CrossRef] [Green Version]
  204. Schmidt, F.; Levin, J.; Kamp, F.; Kretzschmar, H.; Giese, A.; Botzel, K. Single-channel electrophysiology reveals a distinct and uniform pore complex formed by alpha-synuclein oligomers in lipid membranes. PLoS ONE 2012, 7, e42545. [Google Scholar] [CrossRef] [Green Version]
  205. Mironov, S.L. alpha-Synuclein forms non-selective cation channels and stimulates ATP-sensitive potassium channels in hippocampal neurons. J. Physiol. 2015, 593, 145–159. [Google Scholar] [CrossRef] [Green Version]
  206. Shrivastava, A.N.; Redeker, V.; Fritz, N.; Pieri, L.; Almeida, L.G.; Spolidoro, M.; Liebmann, T.; Bousset, L.; Renner, M.; Lena, C.; et al. Alpha-synuclein assemblies sequester neuronal alpha3-Na+/K+-ATPase and impair Na+ gradient. EMBO J. 2015, 34, 2408–2423. [Google Scholar] [CrossRef] [Green Version]
  207. Ronzitti, G.; Bucci, G.; Emanuele, M.; Leo, D.; Sotnikova, T.D.; Mus, L.V.; Soubrane, C.H.; Dallas, M.L.; Thalhammer, A.; Cingolani, L.A.; et al. Exogenous alpha-synuclein decreases raft partitioning of Cav2.2 channels inducing dopamine release. J. Neurosci. 2014, 34, 10603–10615. [Google Scholar] [CrossRef] [PubMed]
  208. Ritz, B.; Rhodes, S.L.; Qian, L.; Schernhammer, E.; Olsen, J.H.; Friis, S. L-type calcium channel blockers and Parkinson disease in Denmark. Ann. Neurol. 2010, 67, 600–606. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Pasternak, B.; Svanstrom, H.; Nielsen, N.M.; Fugger, L.; Melbye, M.; Hviid, A. Use of calcium channel blockers and Parkinson’s disease. Am. J. Epidemiol. 2012, 175, 627–635. [Google Scholar] [CrossRef] [PubMed]
  210. Chan, C.S.; Guzman, J.N.; Ilijic, E.; Mercer, J.N.; Rick, C.; Tkatch, T.; Meredith, G.E.; Surmeier, D.J. ‘Rejuvenation’ protects neurons in mouse models of Parkinson’s disease. Nature 2007, 447, 1081–1086. [Google Scholar] [CrossRef]
  211. Singh, A.; Verma, P.; Balaji, G.; Samantaray, S.; Mohanakumar, K.P. Nimodipine, an L-type calcium channel blocker attenuates mitochondrial dysfunctions to protect against 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine-induced Parkinsonism in mice. Neurochem. Int. 2016, 99, 221–232. [Google Scholar] [CrossRef] [PubMed]
  212. McGranahan, T.M.; Patzlaff, N.E.; Grady, S.R.; Heinemann, S.F.; Booker, T.K. alpha4beta2 nicotinic acetylcholine receptors on dopaminergic neurons mediate nicotine reward and anxiety relief. J. Neurosci. 2011, 31, 10891–10902. [Google Scholar] [CrossRef] [Green Version]
  213. Swant, J.; Goodwin, J.S.; North, A.; Ali, A.A.; Gamble-George, J.; Chirwa, S.; Khoshbouei, H. alpha-Synuclein stimulates a dopamine transporter-dependent chloride current and modulates the activity of the transporter. J. Biol. Chem. 2011, 286, 43933–43943. [Google Scholar] [CrossRef] [Green Version]
  214. Butler, B.; Saha, K.; Rana, T.; Becker, J.P.; Sambo, D.; Davari, P.; Goodwin, J.S.; Khoshbouei, H. Dopamine transporter activity is modulated by alpha-synuclein. J. Biol. Chem. 2015, 290, 29542–29554. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Sidhu, A.; Wersinger, C.; Vernier, P. alpha-Synuclein regulation of the dopaminergic transporter: A possible role in the pathogenesis of Parkinson’s disease. FEBS Lett. 2004, 565, 1–5. [Google Scholar] [CrossRef]
  216. Dean, E.D.; Li, Y.; Torres, G.E.; Miller, G.W. Identification of a novel interaction between α-synuclein and VMAT2. FASEB J. 2008, 22, 715.6. [Google Scholar]
  217. Guo, J.T.; Chen, A.Q.; Kong, Q.; Zhu, H.; Ma, C.M.; Qin, C. Inhibition of vesicular monoamine transporter-2 activity in alpha-synuclein stably transfected SH-SY5Y cells. Cell Mol. Neurobiol. 2008, 28, 35–47. [Google Scholar] [CrossRef] [PubMed]
  218. Perez, R.G.; Waymire, J.C.; Lin, E.; Liu, J.J.; Guo, F.; Zigmond, M.J. A role for alpha-synuclein in the regulation of dopamine biosynthesis. J. Neurosci. 2002, 22, 3090–3099. [Google Scholar] [CrossRef]
  219. Baptista, M.J.; O’Farrell, C.; Daya, S.; Ahmad, R.; Miller, D.W.; Hardy, J.; Farrer, M.J.; Cookson, M.R. Co-ordinate transcriptional regulation of dopamine synthesis genes by alpha-synuclein in human neuroblastoma cell lines. J. Neurochem. 2003, 85, 957–968. [Google Scholar] [CrossRef] [PubMed]
  220. Yu, S.; Zuo, X.; Li, Y.; Zhang, C.; Zhou, M.; Zhang, Y.A.; Ueda, K.; Chan, P. Inhibition of tyrosine hydroxylase expression in alpha-synuclein-transfected dopaminergic neuronal cells. Neurosci. Lett. 2004, 367, 34–39. [Google Scholar] [CrossRef]
  221. Li, Y.H.; Gao, N.; Ye, Y.W.; Li, X.; Yu, S.; Yang, H.; Ueda, K.; Chan, P. Alpha-synuclein functions as a negative regulator for expression of tyrosine hydroxylase. Acta Neurol. Belg. 2011, 111, 130–135. [Google Scholar] [PubMed]
  222. Liu, D.; Jin, L.; Wang, H.; Zhao, H.; Zhao, C.; Duan, C.; Lu, L.; Wu, B.; Yu, S.; Chan, P.; et al. Silencing alpha-synuclein gene expression enhances tyrosine hydroxylase activity in MN9D cells. Neurochem. Res. 2008, 33, 1401–1409. [Google Scholar] [CrossRef] [Green Version]
  223. Peng, X.; Tehranian, R.; Dietrich, P.; Stefanis, L.; Perez, R.G. Alpha-synuclein activation of protein phosphatase 2A reduces tyrosine hydroxylase phosphorylation in dopaminergic cells. J. Cell Sci. 2005, 118, 3523–3530. [Google Scholar] [CrossRef] [Green Version]
  224. Alerte, T.N.; Akinfolarin, A.A.; Friedrich, E.E.; Mader, S.A.; Hong, C.S.; Perez, R.G. Alpha-synuclein aggregation alters tyrosine hydroxylase phosphorylation and immunoreactivity: Lessons from viral transduction of knockout mice. Neurosci. Lett. 2008, 435, 24–29. [Google Scholar] [CrossRef] [Green Version]
  225. Lou, H.; Montoya, S.E.; Alerte, T.N.; Wang, J.; Wu, J.; Peng, X.; Hong, C.S.; Friedrich, E.E.; Mader, S.A.; Pedersen, C.J.; et al. Serine 129 phosphorylation reduces the ability of alpha-synuclein to regulate tyrosine hydroxylase and protein phosphatase 2A in vitro and in vivo. J. Biol. Chem. 2010, 285, 17648–17661. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Chu, Y.; Kordower, J.H. Age-associated increases of alpha-synuclein in monkeys and humans are associated with nigrostriatal dopamine depletion: Is this the target for Parkinson’s disease? Neurobiol. Dis. 2007, 25, 134–149. [Google Scholar] [CrossRef] [PubMed]
  227. Chadchankar, H.; Ihalainen, J.; Tanila, H.; Yavich, L. Decreased reuptake of dopamine in the dorsal striatum in the absence of alpha-synuclein. Brain Res. 2011, 1382, 37–44. [Google Scholar] [CrossRef]
  228. Al-Wandi, A.; Ninkina, N.; Millership, S.; Williamson, S.J.; Jones, P.A.; Buchman, V.L. Absence of alpha-synuclein affects dopamine metabolism and synaptic markers in the striatum of aging mice. Neurobiol. Aging 2010, 31, 796–804. [Google Scholar] [CrossRef] [Green Version]
  229. Robertson, D.C.; Schmidt, O.; Ninkina, N.; Jones, P.A.; Sharkey, J.; Buchman, V.L. Developmental loss and resistance to MPTP toxicity of dopaminergic neurones in substantia nigra pars compacta of gamma-synuclein, alpha-synuclein and double alpha/gamma-synuclein null mutant mice. J. Neurochem. 2004, 89, 1126–1136. [Google Scholar] [CrossRef] [PubMed]
  230. Shen, J.; Du, T.; Wang, X.; Duan, C.; Gao, G.; Zhang, J.; Lu, L.; Yang, H. alpha-Synuclein amino terminus regulates mitochondrial membrane permeability. Brain Res. 2014, 1591, 14–26. [Google Scholar] [CrossRef] [PubMed]
  231. Nakamura, K.; Nemani, V.M.; Azarbal, F.; Skibinski, G.; Levy, J.M.; Egami, K.; Munishkina, L.; Zhang, J.; Gardner, B.; Wakabayashi, J.; et al. Direct membrane association drives mitochondrial fission by the Parkinson disease-associated protein alpha-synuclein. J. Biol. Chem. 2011, 286, 20710–20726. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  232. Robotta, M.; Gerding, H.R.; Vogel, A.; Hauser, K.; Schildknecht, S.; Karreman, C.; Leist, M.; Subramaniam, V.; Drescher, M. Alpha-synuclein binds to the inner membrane of mitochondria in an alpha-helical conformation. ChemBioChem 2014, 15, 2499–2502. [Google Scholar] [CrossRef] [PubMed]
  233. Faustini, G.; Marchesan, E.; Zonta, L.; Bono, F.; Bottani, E.; Longhena, F.; Ziviani, E.; Valerio, A.; Bellucci, A. Alpha-synuclein preserves mitochondrial fusion and function in neuronal cells. Oxid. Med. Cell Longev. 2019, 2019, 4246350. [Google Scholar] [CrossRef] [Green Version]
  234. Ellis, C.E.; Murphy, E.J.; Mitchell, D.C.; Golovko, M.Y.; Scaglia, F.; Barcelo-Coblijn, G.C.; Nussbaum, R.L. Mitochondrial lipid abnormality and electron transport chain impairment in mice lacking alpha-synuclein. Mol. Cell Biol. 2005, 25, 10190–10201. [Google Scholar] [CrossRef] [Green Version]
  235. Ludtmann, M.H.; Angelova, P.R.; Ninkina, N.N.; Gandhi, S.; Buchman, V.L.; Abramov, A.Y. Monomeric alpha-synuclein exerts a physiological role on brain ATP synthase. J. Neurosci. 2016, 36, 10510–10521. [Google Scholar] [CrossRef]
  236. Hsu, L.J.; Sagara, Y.; Arroyo, A.; Rockenstein, E.; Sisk, A.; Mallory, M.; Wong, J.; Takenouchi, T.; Hashimoto, M.; Masliah, E. Alpha-synuclein promotes mitochondrial deficit and oxidative stress. Am. J. Pathol. 2000, 157, 401–410. [Google Scholar] [CrossRef]
  237. Parihar, M.S.; Parihar, A.; Fujita, M.; Hashimoto, M.; Ghafourifar, P. Mitochondrial association of alpha-synuclein causes oxidative stress. Cell Mol. Life Sci. 2008, 65, 1272–1284. [Google Scholar] [CrossRef]
  238. Hu, D.; Sun, X.; Liao, X.; Zhang, X.; Zarabi, S.; Schimmer, A.; Hong, Y.; Ford, C.; Luo, Y.; Qi, X. Alpha-synuclein suppresses mitochondrial protease ClpP to trigger mitochondrial oxidative damage and neurotoxicity. Acta Neuropathol. 2019, 137, 939–960. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  239. Di Maio, R.; Barrett, P.J.; Hoffman, E.K.; Barrett, C.W.; Zharikov, A.; Borah, A.; Hu, X.; McCoy, J.; Chu, C.T.; Burton, E.A.; et al. Alpha-synuclein binds to TOM20 and inhibits mitochondrial protein import in Parkinson’s disease. Sci. Transl. Med. 2016, 8, 342ra78. [Google Scholar] [CrossRef] [Green Version]
  240. Devi, L.; Raghavendran, V.; Prabhu, B.M.; Avadhani, N.G.; Anandatheerthavarada, H.K. Mitochondrial import and accumulation of alpha-synuclein impair complex I in human dopaminergic neuronal cultures and Parkinson disease brain. J. Biol. Chem. 2008, 283, 9089–9100. [Google Scholar] [CrossRef] [Green Version]
  241. Bir, A.; Sen, O.; Anand, S.; Khemka, V.K.; Banerjee, P.; Cappai, R.; Sahoo, A.; Chakrabarti, S. alpha-Synuclein-induced mitochondrial dysfunction in isolated preparation and intact cells: Implications in the pathogenesis of Parkinson’s disease. J. Neurochem. 2014, 131, 868–877. [Google Scholar] [CrossRef]
  242. Perfeito, R.; Lazaro, D.F.; Outeiro, T.F.; Rego, A.C. Linking alpha-synuclein phosphorylation to reactive oxygen species formation and mitochondrial dysfunction in SH-SY5Y cells. Mol. Cell Neurosci. 2014, 62, 51–59. [Google Scholar] [CrossRef] [PubMed]
  243. Subramaniam, S.R.; Vergnes, L.; Franich, N.R.; Reue, K.; Chesselet, M.F. Region specific mitochondrial impairment in mice with widespread overexpression of alpha-synuclein. Neurobiol. Dis. 2014, 70, 204–213. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Martin, L.J.; Pan, Y.; Price, A.C.; Sterling, W.; Copeland, N.G.; Jenkins, N.A.; Price, D.L.; Lee, M.K. Parkinson’s disease alpha-synuclein transgenic mice develop neuronal mitochondrial degeneration and cell death. J. Neurosci. 2006, 26, 41–50. [Google Scholar] [CrossRef] [Green Version]
  245. Fornai, F.; Schluter, O.M.; Lenzi, P.; Gesi, M.; Ruffoli, R.; Ferrucci, M.; Lazzeri, G.; Busceti, C.L.; Pontarelli, F.; Battaglia, G.; et al. Parkinson-like syndrome induced by continuous MPTP infusion: Convergent roles of the ubiquitin-proteasome system and alpha-synuclein. Proc. Natl. Acad. Sci. USA 2005, 102, 3413–3418. [Google Scholar] [CrossRef] [Green Version]
  246. Manning-Bog, A.B.; McCormack, A.L.; Li, J.; Uversky, V.N.; Fink, A.L.; Di Monte, D.A. The herbicide paraquat causes up-regulation and aggregation of alpha-synuclein in mice: Paraquat and alpha-synuclein. J. Biol. Chem. 2002, 277, 1641–1644. [Google Scholar] [CrossRef] [Green Version]
  247. Radad, K.; Al-Shraim, M.; Al-Emam, A.; Wang, F.; Kranner, B.; Rausch, W.D.; Moldzio, R. Rotenone: From modelling to implication in Parkinson’s disease. Folia Neuropathol. 2019, 57, 317–326. [Google Scholar] [CrossRef]
  248. Vila, M.; Vukosavic, S.; Jackson-Lewis, V.; Neystat, M.; Jakowec, M.; Przedborski, S. Alpha-synuclein up-regulation in substantia nigra dopaminergic neurons following administration of the parkinsonian toxin MPTP. J. Neurochem. 2000, 74, 721–729. [Google Scholar] [CrossRef] [PubMed]
  249. Chinta, S.J.; Mallajosyula, J.K.; Rane, A.; Andersen, J.K. Mitochondrial alpha-synuclein accumulation impairs complex I function in dopaminergic neurons and results in increased mitophagy in vivo. Neurosci. Lett. 2010, 486, 235–239. [Google Scholar] [CrossRef] [Green Version]
  250. Loeb, V.; Yakunin, E.; Saada, A.; Sharon, R. The transgenic overexpression of alpha-synuclein and not its related pathology associates with complex I inhibition. J. Biol. Chem. 2010, 285, 7334–7343. [Google Scholar] [CrossRef] [Green Version]
  251. Luth, E.S.; Stavrovskaya, I.G.; Bartels, T.; Kristal, B.S.; Selkoe, D.J. Soluble, prefibrillar alpha-synuclein oligomers promote complex I-dependent, Ca2+-induced mitochondrial dysfunction. J. Biol. Chem. 2014, 289, 21490–21507. [Google Scholar] [CrossRef] [Green Version]
  252. Paillusson, S.; Gomez-Suaga, P.; Stoica, R.; Little, D.; Gissen, P.; Devine, M.J.; Noble, W.; Hanger, D.P.; Miller, C.C.J. alpha-Synuclein binds to the ER-mitochondria tethering protein VAPB to disrupt Ca(2+) homeostasis and mitochondrial ATP production. Acta Neuropathol. 2017, 134, 129–149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  253. Chen, L.; Xie, Z.; Turkson, S.; Zhuang, X. A53T human alpha-synuclein overexpression in transgenic mice induces pervasive mitochondria macroautophagy defects preceding dopamine neuron degeneration. J. Neurosci. 2015, 35, 890–905. [Google Scholar] [CrossRef]
  254. Poon, H.F.; Frasier, M.; Shreve, N.; Calabrese, V.; Wolozin, B.; Butterfield, D.A. Mitochondrial associated metabolic proteins are selectively oxidized in A30P alpha-synuclein transgenic mice—A model of familial Parkinson’s disease. Neurobiol. Dis. 2005, 18, 492–498. [Google Scholar] [CrossRef] [PubMed]
  255. Kamp, F.; Exner, N.; Lutz, A.K.; Wender, N.; Hegermann, J.; Brunner, B.; Nuscher, B.; Bartels, T.; Giese, A.; Beyer, K.; et al. Inhibition of mitochondrial fusion by alpha-synuclein is rescued by PINK1, Parkin and DJ-1. EMBO J. 2010, 29, 3571–3589. [Google Scholar] [CrossRef] [Green Version]
  256. Betarbet, R.; Sherer, T.B.; MacKenzie, G.; Garcia-Osuna, M.; Panov, A.V.; Greenamyre, J.T. Chronic systemic pesticide exposure reproduces features of Parkinson’s disease. Nat. Neurosci. 2000, 3, 1301–1306. [Google Scholar] [CrossRef]
  257. Nistico, R.; Mehdawy, B.; Piccirilli, S.; Mercuri, N. Paraquat- and rotenone-induced models of Parkinson’s disease. Int. J. Immunopathol. Pharmacol. 2011, 24, 313–322. [Google Scholar] [CrossRef] [Green Version]
  258. Tieu, K. A guide to neurotoxic animal models of Parkinson’s disease. Cold Spring Harb. Perspect. Med. 2011, 1, a009316. [Google Scholar] [CrossRef]
  259. Lee, H.J.; Shin, S.Y.; Choi, C.; Lee, Y.H.; Lee, S.J. Formation and removal of alpha-synuclein aggregates in cells exposed to mitochondrial inhibitors. J. Biol. Chem. 2002, 277, 5411–5417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  260. Kowall, N.W.; Hantraye, P.; Brouillet, E.; Beal, M.F.; McKee, A.C.; Ferrante, R.J. MPTP induces alpha-synuclein aggregation in the substantia nigra of baboons. Neuroreport 2000, 11, 211–213. [Google Scholar] [CrossRef] [PubMed]
  261. Przedborski, S.; Chen, Q.; Vila, M.; Giasson, B.I.; Djaldatti, R.; Vukosavic, S.; Souza, J.M.; Jackson-Lewis, V.; Lee, V.M.; Ischiropoulos, H. Oxidative post-translational modifications of alpha-synuclein in the 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) mouse model of Parkinson’s disease. J. Neurochem. 2001, 76, 637–640. [Google Scholar] [CrossRef] [PubMed]
  262. Hu, S.; Hu, M.; Liu, J.; Zhang, B.; Zhang, Z.; Zhou, F.H.; Wang, L.; Dong, J. Phosphorylation of tau and alpha-synuclein induced neurodegeneration in MPTP mouse model of Parkinson’s disease. Neuropsychiatry Dis. Treat. 2020, 16, 651–663. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  263. Lee, S.; Oh, S.T.; Jeong, H.J.; Pak, S.C.; Park, H.J.; Kim, J.; Cho, H.S.; Jeon, S. MPTP-induced vulnerability of dopamine neurons in A53T alpha-synuclein overexpressed mice with the potential involvement of DJ-1 downregulation. Korean J. Physiol. Pharmacol. 2017, 21, 625–632. [Google Scholar] [CrossRef] [Green Version]
  264. Bazzu, G.; Calia, G.; Puggioni, G.; Spissu, Y.; Rocchitta, G.; Debetto, P.; Grigoletto, J.; Zusso, M.; Migheli, R.; Serra, P.A.; et al. alpha-Synuclein- and MPTP-generated rodent models of Parkinson’s disease and the study of extracellular striatal dopamine dynamics: A microdialysis approach. CNS Neurol. Disord. Drug Targets 2010, 9, 482–490. [Google Scholar] [CrossRef]
  265. Grunewald, A.; Arns, B.; Seibler, P.; Rakovic, A.; Munchau, A.; Ramirez, A.; Sue, C.M.; Klein, C. ATP13A2 mutations impair mitochondrial function in fibroblasts from patients with Kufor-Rakeb syndrome. Neurobiol. Aging 2012, 33, 1843.e1. [Google Scholar] [CrossRef]
  266. Park, J.S.; Koentjoro, B.; Veivers, D.; Mackay-Sim, A.; Sue, C.M. Parkinson’s disease-associated human ATP13A2 (PARK9) deficiency causes zinc dyshomeostasis and mitochondrial dysfunction. Hum. Mol. Genet. 2014, 23, 2802–2815. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  267. Cali, T.; Ottolini, D.; Negro, A.; Brini, M. alpha-Synuclein controls mitochondrial calcium homeostasis by enhancing endoplasmic reticulum-mitochondria interactions. J. Biol. Chem. 2012, 287, 17914–17929. [Google Scholar] [CrossRef] [Green Version]
  268. Guardia-Laguarta, C.; Area-Gomez, E.; Schon, E.A.; Przedborski, S. A new role for alpha-synuclein in Parkinson’s disease: Alteration of ER-mitochondrial communication. Mov. Disord. 2015, 30, 1026–1033. [Google Scholar] [CrossRef]
  269. Guardia-Laguarta, C.; Area-Gomez, E.; Rub, C.; Liu, Y.; Magrane, J.; Becker, D.; Voos, W.; Schon, E.A.; Przedborski, S. alpha-Synuclein is localized to mitochondria-associated ER membranes. J. Neurosci. 2014, 34, 249–259. [Google Scholar] [CrossRef] [PubMed]
  270. Jacobs, D.; Hoogerheide, D.P.; Rovini, A.; Jiang, Z.; Lee, J.C.; Rostovtseva, T.K.; Bezrukov, S.M. Probing membrane association of alpha-synuclein domains with VDAC nanopore reveals unexpected binding pattern. Sci. Rep. 2019, 9, 4580. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  271. Rostovtseva, T.K.; Gurnev, P.A.; Protchenko, O.; Hoogerheide, D.P.; Yap, T.L.; Philpott, C.C.; Lee, J.C.; Bezrukov, S.M. Alpha-synuclein shows high affinity interaction with voltage-dependent anion channel, suggesting mechanisms of mitochondrial regulation and toxicity in Parkinson disease. J. Biol. Chem. 2015, 290, 18467–18477. [Google Scholar] [CrossRef] [Green Version]
  272. Hoogerheide, D.P.; Gurnev, P.A.; Rostovtseva, T.K.; Bezrukov, S.M. Mechanism of alpha-synuclein translocation through a VDAC nanopore revealed by energy landscape modeling of escape time distributions. Nanoscale 2017, 9, 183–192. [Google Scholar] [CrossRef]
  273. Lu, L.; Zhang, C.; Cai, Q.; Lu, Q.; Duan, C.; Zhu, Y.; Yang, H. Voltage-dependent anion channel involved in the alpha-synuclein-induced dopaminergic neuron toxicity in rats. Acta Biochim. Biophys. Sin. 2013, 45, 170–178. [Google Scholar] [CrossRef] [Green Version]
  274. Shoshan-Barmatz, V.; Zalk, R.; Gincel, D.; Vardi, N. Subcellular localization of VDAC in mitochondria and ER in the cerebellum. Biochim. Biophys. Acta 2004, 1657, 105–114. [Google Scholar] [CrossRef] [Green Version]
  275. Shoshan-Barmatz, V.; Israelson, A. The voltage-dependent anion channel in endoplasmic/sarcoplasmic reticulum: Characterization, modulation and possible function. J. Membr. Biol. 2005, 204, 57–66. [Google Scholar] [CrossRef] [PubMed]
  276. De Stefani, D.; Bononi, A.; Romagnoli, A.; Messina, A.; De Pinto, V.; Pinton, P.; Rizzuto, R. VDAC1 selectively transfers apoptotic Ca2+ signals to mitochondria. Cell Death Differ. 2012, 19, 267–273. [Google Scholar] [CrossRef] [Green Version]
  277. Chu, Y.; Goldman, J.G.; Kelly, L.; He, Y.; Waliczek, T.; Kordower, J.H. Abnormal alpha-synuclein reduces nigral voltage-dependent anion channel 1 in sporadic and experimental Parkinson’s disease. Neurobiol. Dis. 2014, 69, 1–14. [Google Scholar] [CrossRef] [PubMed]
  278. Ludtmann, M.H.R.; Angelova, P.R.; Horrocks, M.H.; Choi, M.L.; Rodrigues, M.; Baev, A.Y.; Berezhnov, A.V.; Yao, Z.; Little, D.; Banushi, B.; et al. Alpha-synuclein oligomers interact with ATP synthase and open the permeability transition pore in Parkinson’s disease. Nat. Commun. 2018, 9, 2293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  279. Martin, L.J.; Semenkow, S.; Hanaford, A.; Wong, M. Mitochondrial permeability transition pore regulates Parkinson’s disease development in mutant alpha-synuclein transgenic mice. Neurobiol. Aging 2014, 35, 1132–1152. [Google Scholar] [CrossRef] [Green Version]
  280. Heman-Ackah, S.M.; Manzano, R.; Hoozemans, J.J.M.; Scheper, W.; Flynn, R.; Haerty, W.; Cowley, S.A.; Bassett, A.R.; Wood, M.J.A. Alpha-synuclein induces the unfolded protein response in Parkinson’s disease SNCA triplication iPSC-derived neurons. Hum. Mol. Genet. 2017, 26, 4441–4450. [Google Scholar] [CrossRef] [Green Version]
  281. Lee, A.S. The ER chaperone and signaling regulator GRP78/BiP as a monitor of endoplasmic reticulum stress. Methods 2005, 35, 373–381. [Google Scholar] [CrossRef]
  282. Wang, M.; Wey, S.; Zhang, Y.; Ye, R.; Lee, A.S. Role of the unfolded protein response regulator GRP78/BiP in development, cancer, and neurological disorders. Antioxid. Redox. Signal. 2009, 11, 2307–2316. [Google Scholar] [CrossRef] [PubMed]
  283. Bellucci, A.; Navarria, L.; Zaltieri, M.; Falarti, E.; Bodei, S.; Sigala, S.; Battistin, L.; Spillantini, M.; Missale, C.; Spano, P. Induction of the unfolded protein response by alpha-synuclein in experimental models of Parkinson’s disease. J. Neurochem. 2011, 116, 588–605. [Google Scholar] [CrossRef]
  284. Colla, E.; Coune, P.; Liu, Y.; Pletnikova, O.; Troncoso, J.C.; Iwatsubo, T.; Schneider, B.L.; Lee, M.K. Endoplasmic reticulum stress is important for the manifestations of alpha-synucleinopathy in vivo. J. Neurosci. 2012, 32, 3306–3320. [Google Scholar] [CrossRef] [PubMed]
  285. Colla, E.; Jensen, P.H.; Pletnikova, O.; Troncoso, J.C.; Glabe, C.; Lee, M.K. Accumulation of toxic alpha-synuclein oligomer within endoplasmic reticulum occurs in alpha-synucleinopathy in vivo. J. Neurosci. 2012, 32, 3301–3305. [Google Scholar] [CrossRef] [Green Version]
  286. Baek, J.H.; Whitfield, D.; Howlett, D.; Francis, P.; Bereczki, E.; Ballard, C.; Hortobagyi, T.; Attems, J.; Aarsland, D. Unfolded protein response is activated in Lewy body dementias. Neuropathol. Appl. Neurobiol. 2016, 42, 352–365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  287. Yagi-Utsumi, M.; Satoh, T.; Kato, K. Structural basis of redox-dependent substrate binding of protein disulfide isomerase. Sci. Rep. 2015, 5, 13909. [Google Scholar] [CrossRef] [Green Version]
  288. Honjo, Y.; Ito, H.; Horibe, T.; Takahashi, R.; Kawakami, K. Protein disulfide isomerase immunopositive glial cytoplasmic inclusions in patients with multiple system atrophy. Int. J. Neurosci. 2011, 121, 543–550. [Google Scholar] [CrossRef]
  289. Conn, K.J.; Gao, W.; McKee, A.; Lan, M.S.; Ullman, M.D.; Eisenhauer, P.B.; Fine, R.E.; Wells, J.M. Identification of the protein disulfide isomerase family member PDIp in experimental Parkinson’s disease and Lewy body pathology. Brain Res. 2004, 1022, 164–172. [Google Scholar] [CrossRef]
  290. Jin, J.; Li, G.J.; Davis, J.; Zhu, D.; Wang, Y.; Pan, C.; Zhang, J. Identification of novel proteins associated with both alpha-synuclein and DJ-1. Mol. Cell Proteom. 2007, 6, 845–859. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  291. Koch, G.; Smith, M.; Macer, D.; Webster, P.; Mortara, R. Endoplasmic reticulum contains a common, abundant calcium-binding glycoprotein, endoplasmin. J. Cell Sci. 1986, 86, 217–232. [Google Scholar] [CrossRef]
  292. Labrador-Garrido, A.; Cejudo-Guillen, M.; Daturpalli, S.; Leal, M.M.; Klippstein, R.; De Genst, E.J.; Villadiego, J.; Toledo-Aral, J.J.; Dobson, C.M.; Jackson, S.E.; et al. Chaperome screening leads to identification of Grp94/Gp96 and FKBP4/52 as modulators of the alpha-synuclein-elicited immune response. FASEB J. 2016, 30, 564–577. [Google Scholar] [CrossRef] [Green Version]
  293. Lee, A.S. Mammalian stress response: Induction of the glucose-regulated protein family. Curr. Opin. Cell Biol. 1992, 4, 267–273. [Google Scholar] [CrossRef]
  294. Cooper, A.A.; Gitler, A.D.; Cashikar, A.; Haynes, C.M.; Hill, K.J.; Bhullar, B.; Liu, K.; Xu, K.; Strathearn, K.E.; Liu, F.; et al. Alpha-synuclein blocks ER-Golgi traffic and Rab1 rescues neuron loss in Parkinson’s models. Science 2006, 313, 324–328. [Google Scholar] [CrossRef] [Green Version]
  295. Thayanidhi, N.; Helm, J.R.; Nycz, D.C.; Bentley, M.; Liang, Y.; Hay, J.C. Alpha-synuclein delays endoplasmic reticulum (ER)-to-Golgi transport in mammalian cells by antagonizing ER/Golgi SNAREs. Mol. Biol. Cell 2010, 21, 1850–1863. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  296. Gitler, A.D.; Bevis, B.J.; Shorter, J.; Strathearn, K.E.; Hamamichi, S.; Su, L.J.; Caldwell, K.A.; Caldwell, G.A.; Rochet, J.C.; McCaffery, J.M.; et al. The Parkinson’s disease protein alpha-synuclein disrupts cellular Rab homeostasis. Proc. Natl. Acad. Sci. USA 2008, 105, 145–150. [Google Scholar] [CrossRef] [Green Version]
  297. Dalfo, E.; Gomez-Isla, T.; Rosa, J.L.; Nieto Bodelon, M.; Cuadrado Tejedor, M.; Barrachina, M.; Ambrosio, S.; Ferrer, I. Abnormal alpha-synuclein interactions with Rab proteins in alpha-synuclein A30P transgenic mice. J. Neuropathol. Exp. Neurol. 2004, 63, 302–313. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  298. Yin, G.; Lopes da Fonseca, T.; Eisbach, S.E.; Anduaga, A.M.; Breda, C.; Orcellet, M.L.; Szego, E.M.; Guerreiro, P.; Lazaro, D.F.; Braus, G.H.; et al. alpha-Synuclein interacts with the switch region of Rab8a in a Ser129 phosphorylation-dependent manner. Neurobiol. Dis. 2014, 70, 149–161. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  299. Breda, C.; Nugent, M.L.; Estranero, J.G.; Kyriacou, C.P.; Outeiro, T.F.; Steinert, J.R.; Giorgini, F. Rab11 modulates alpha-synuclein-mediated defects in synaptic transmission and behaviour. Hum. Mol. Genet. 2015, 24, 1077–1091. [Google Scholar] [CrossRef] [Green Version]
  300. Fujita, Y.; Ohama, E.; Takatama, M.; Al-Sarraj, S.; Okamoto, K. Fragmentation of Golgi apparatus of nigral neurons with alpha-synuclein-positive inclusions in patients with Parkinson’s disease. Acta Neuropathol. 2006, 112, 261–265. [Google Scholar] [CrossRef] [PubMed]
  301. Lee, H.J.; Khoshaghideh, F.; Lee, S.; Lee, S.J. Impairment of microtubule-dependent trafficking by overexpression of alpha-synuclein. Eur. J. Neurosci. 2006, 24, 3153–3162. [Google Scholar] [CrossRef]
  302. Gosavi, N.; Lee, H.J.; Lee, J.S.; Patel, S.; Lee, S.J. Golgi fragmentation occurs in the cells with prefibrillar alpha-synuclein aggregates and precedes the formation of fibrillar inclusion. J. Biol. Chem. 2002, 277, 48984–48992. [Google Scholar] [CrossRef] [Green Version]
  303. Betzer, C.; Lassen, L.B.; Olsen, A.; Kofoed, R.H.; Reimer, L.; Gregersen, E.; Zheng, J.; Cali, T.; Gai, W.P.; Chen, T.; et al. Alpha-synuclein aggregates activate calcium pump SERCA leading to calcium dysregulation. EMBO Rep. 2018, 19. [Google Scholar] [CrossRef] [PubMed]
  304. Wuytack, F.; Raeymaekers, L.; Missiaen, L. PMR1/SPCA Ca2+ pumps and the role of the Golgi apparatus as a Ca2+ store. Pflugers Arch. 2003, 446, 148–153. [Google Scholar] [CrossRef]
  305. Buttner, S.; Faes, L.; Reichelt, W.N.; Broeskamp, F.; Habernig, L.; Benke, S.; Kourtis, N.; Ruli, D.; Carmona-Gutierrez, D.; Eisenberg, T.; et al. The Ca2+/Mn2+ ion-pump PMR1 links elevation of cytosolic Ca(2+) levels to alpha-synuclein toxicity in Parkinson’s disease models. Cell Death Differ. 2013, 20, 465–477. [Google Scholar] [CrossRef] [Green Version]
  306. Nikoletopoulou, V.; Tavernarakis, N. The PMR1 pump in alpha-synuclein toxicity and neurodegeneration. Neurosci. Lett. 2018, 663, 66–71. [Google Scholar] [CrossRef] [PubMed]
  307. Goers, J.; Manning-Bog, A.B.; McCormack, A.L.; Millett, I.S.; Doniach, S.; Di Monte, D.A.; Uversky, V.N.; Fink, A.L. Nuclear localization of alpha-synuclein and its interaction with histones. Biochemistry 2003, 42, 8465–8471. [Google Scholar] [CrossRef]
  308. Mori, F.; Tanji, K.; Yoshimoto, M.; Takahashi, H.; Wakabayashi, K. Immunohistochemical comparison of alpha- and beta-synuclein in adult rat central nervous system. Brain Res. 2002, 941, 118–126. [Google Scholar] [CrossRef]
  309. Yu, S.; Li, X.; Liu, G.; Han, J.; Zhang, C.; Li, Y.; Xu, S.; Liu, C.; Gao, Y.; Yang, H.; et al. Extensive nuclear localization of alpha-synuclein in normal rat brain neurons revealed by a novel monoclonal antibody. Neuroscience 2007, 145, 539–555. [Google Scholar] [CrossRef]
  310. Huang, Z.; Xu, Z.; Wu, Y.; Zhou, Y. Determining nuclear localization of alpha-synuclein in synuclein mouse brains. Neuroscience 2011, 199, 318–332. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  311. Kontopoulos, E.; Parvin, J.D.; Feany, M.B. Alpha-synuclein acts in the nucleus to inhibit histone acetylation and promote neurotoxicity. Hum. Mol. Genet. 2006, 15, 3012–3023. [Google Scholar] [CrossRef] [Green Version]
  312. Ma, K.L.; Song, L.K.; Yuan, Y.H.; Zhang, Y.; Han, N.; Gao, K.; Chen, N.H. The nuclear accumulation of alpha-synuclein is mediated by importin alpha and promotes neurotoxicity by accelerating the cell cycle. Neuropharmacology 2014, 82, 132–142. [Google Scholar] [CrossRef] [PubMed]
  313. Desplats, P.; Spencer, B.; Coffee, E.; Patel, P.; Michael, S.; Patrick, C.; Adame, A.; Rockenstein, E.; Masliah, E. Alpha-synuclein sequesters Dnmt1 from the nucleus: A novel mechanism for epigenetic alterations in Lewy body diseases. J. Biol. Chem. 2011, 286, 9031–9037. [Google Scholar] [CrossRef] [Green Version]
  314. Outeiro, T.F.; Kontopoulos, E.; Altmann, S.M.; Kufareva, I.; Strathearn, K.E.; Amore, A.M.; Volk, C.B.; Maxwell, M.M.; Rochet, J.C.; McLean, P.J.; et al. Sirtuin 2 inhibitors rescue alpha-synuclein-mediated toxicity in models of Parkinson’s disease. Science 2007, 317, 516–519. [Google Scholar] [CrossRef] [PubMed]
  315. Davidi, D.; Schechter, M.; Elhadi, S.A.; Matatov, A.; Nathanson, L.; Sharon, R. Alpha-synuclein translocates to the nucleus to activate retinoic-acid-dependent gene transcription. iScience 2020, 23, 100910. [Google Scholar] [CrossRef] [Green Version]
  316. Schell, H.; Hasegawa, T.; Neumann, M.; Kahle, P.J. Nuclear and neuritic distribution of serine-129 phosphorylated alpha-synuclein in transgenic mice. Neuroscience 2009, 160, 796–804. [Google Scholar] [CrossRef]
  317. Pinho, R.; Paiva, I.; Jercic, K.G.; Fonseca-Ornelas, L.; Gerhardt, E.; Fahlbusch, C.; Garcia-Esparcia, P.; Kerimoglu, C.; Pavlou, M.A.S.; Villar-Pique, A.; et al. Nuclear localization and phosphorylation modulate pathological effects of alpha-synuclein. Hum. Mol. Genet. 2019, 28, 31–50. [Google Scholar] [CrossRef] [PubMed]
  318. Ryu, S.B.I.; Liew, H. Sumoylated α-synuclein translocates into the nucleus by karyopherin α6. Mol. Cell. Toxicol. 2019, 15, 103–109. [Google Scholar] [CrossRef]
  319. Chen, V.; Moncalvo, M.; Tringali, D.; Tagliafierro, L.; Shriskanda, A.; Ilich, E.; Dong, W.; Kantor, B.; Chiba-Falek, O. The mechanistic role of alpha-synuclein in the nucleus: Impaired nuclear function caused by familial Parkinson’s disease SNCA mutations. Hum. Mol. Genet. 2020, 29, 3107–3121. [Google Scholar] [CrossRef] [PubMed]
  320. Schaser, A.J.; Osterberg, V.R.; Dent, S.E.; Stackhouse, T.L.; Wakeham, C.M.; Boutros, S.W.; Weston, L.J.; Owen, N.; Weissman, T.A.; Luna, E.; et al. Alpha-synuclein is a DNA binding protein that modulates DNA repair with implications for Lewy body disorders. Sci. Rep. 2019, 9, 10919. [Google Scholar] [CrossRef] [Green Version]
  321. Liu, X.; Lee, Y.J.; Liou, L.C.; Ren, Q.; Zhang, Z.; Wang, S.; Witt, S.N. Alpha-synuclein functions in the nucleus to protect against hydroxyurea-induced replication stress in yeast. Hum. Mol. Genet. 2011, 20, 3401–3414. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  322. Bennett, M.C.; Bishop, J.F.; Leng, Y.; Chock, P.B.; Chase, T.N.; Mouradian, M.M. Degradation of alpha-synuclein by proteasome. J. Biol. Chem. 1999, 274, 33855–33858. [Google Scholar] [CrossRef] [Green Version]
  323. Webb, J.L.; Ravikumar, B.; Atkins, J.; Skepper, J.N.; Rubinsztein, D.C. Alpha-synuclein is degraded by both autophagy and the proteasome. J. Biol. Chem. 2003, 278, 25009–25013. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  324. Vogiatzi, T.; Xilouri, M.; Vekrellis, K.; Stefanis, L. Wild type alpha-synuclein is degraded by chaperone-mediated autophagy and macroautophagy in neuronal cells. J. Biol. Chem. 2008, 283, 23542–23556. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  325. Rott, R.; Szargel, R.; Haskin, J.; Bandopadhyay, R.; Lees, A.J.; Shani, V.; Engelender, S. alpha-Synuclein fate is determined by USP9X-regulated monoubiquitination. Proc. Natl. Acad. Sci. USA 2011, 108, 18666–18671. [Google Scholar] [CrossRef] [Green Version]
  326. Haj-Yahya, M.; Fauvet, B.; Herman-Bachinsky, Y.; Hejjaoui, M.; Bavikar, S.N.; Karthikeyan, S.V.; Ciechanover, A.; Lashuel, H.A.; Brik, A. Synthetic polyubiquitinated alpha-Synuclein reveals important insights into the roles of the ubiquitin chain in regulating its pathophysiology. Proc. Natl. Acad. Sci. USA 2013, 110, 17726–17731. [Google Scholar] [CrossRef] [Green Version]
  327. Tofaris, G.K.; Layfield, R.; Spillantini, M.G. alpha-synuclein metabolism and aggregation is linked to ubiquitin-independent degradation by the proteasome. FEBS Lett. 2001, 509, 22–26. [Google Scholar] [CrossRef] [Green Version]
  328. Liu, C.W.; Corboy, M.J.; DeMartino, G.N.; Thomas, P.J. Endoproteolytic activity of the proteasome. Science 2003, 299, 408–411. [Google Scholar] [CrossRef] [Green Version]
  329. Rideout, H.J.; Larsen, K.E.; Sulzer, D.; Stefanis, L. Proteasomal inhibition leads to formation of ubiquitin/alpha-synuclein-immunoreactive inclusions in PC12 cells. J. Neurochem. 2001, 78, 899–908. [Google Scholar] [CrossRef]
  330. Rideout, H.J.; Stefanis, L. Proteasomal inhibition-induced inclusion formation and death in cortical neurons require transcription and ubiquitination. Mol. Cell. Neurosci. 2002, 21, 223–238. [Google Scholar] [CrossRef]
  331. Emmanouilidou, E.; Stefanis, L.; Vekrellis, K. Cell-produced alpha-synuclein oligomers are targeted to, and impair, the 26S proteasome. Neurobiol. Aging 2010, 31, 953–968. [Google Scholar] [CrossRef] [PubMed]
  332. Ebrahimi-Fakhari, D.; Cantuti-Castelvetri, I.; Fan, Z.; Rockenstein, E.; Masliah, E.; Hyman, B.T.; McLean, P.J.; Unni, V.K. Distinct roles in vivo for the ubiquitin-proteasome system and the autophagy-lysosomal pathway in the degradation of alpha-synuclein. J. Neurosci. 2011, 31, 14508–14520. [Google Scholar] [CrossRef] [PubMed]
  333. Xilouri, M.; Brekk, O.R.; Stefanis, L. alpha-Synuclein and protein degradation systems: A reciprocal relationship. Mol. Neurobiol. 2013, 47, 537–551. [Google Scholar] [CrossRef]
  334. Cuervo, A.M.; Stefanis, L.; Fredenburg, R.; Lansbury, P.T.; Sulzer, D. Impaired degradation of mutant alpha-synuclein by chaperone-mediated autophagy. Science 2004, 305, 1292–1295. [Google Scholar] [CrossRef] [PubMed]
  335. Lee, H.J.; Khoshaghideh, F.; Patel, S.; Lee, S.J. Clearance of alpha-synuclein oligomeric intermediates via the lysosomal degradation pathway. J. Neurosci. 2004, 24, 1888–1896. [Google Scholar] [CrossRef]
  336. Mak, S.K.; McCormack, A.L.; Manning-Bog, A.B.; Cuervo, A.M.; Di Monte, D.A. Lysosomal degradation of alpha-synuclein in vivo. J. Biol. Chem. 2010, 285, 13621–13629. [Google Scholar] [CrossRef] [Green Version]
  337. Xilouri, M.; Brekk, O.R.; Landeck, N.; Pitychoutis, P.M.; Papasilekas, T.; Papadopoulou-Daifoti, Z.; Kirik, D.; Stefanis, L. Boosting chaperone-mediated autophagy in vivo mitigates alpha-synuclein-induced neurodegeneration. Brain 2013, 136, 2130–2146. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  338. Issa, A.R.; Sun, J.; Petitgas, C.; Mesquita, A.; Dulac, A.; Robin, M.; Mollereau, B.; Jenny, A.; Cherif-Zahar, B.; Birman, S. The lysosomal membrane protein LAMP2A promotes autophagic flux and prevents SNCA-induced Parkinson disease-like symptoms in the Drosophila brain. Autophagy 2018, 14, 1898–1910. [Google Scholar] [CrossRef] [Green Version]
  339. Xilouri, M.; Brekk, O.R.; Polissidis, A.; Chrysanthou-Piterou, M.; Kloukina, I.; Stefanis, L. Impairment of chaperone-mediated autophagy induces dopaminergic neurodegeneration in rats. Autophagy 2016, 12, 2230–2247. [Google Scholar] [CrossRef] [Green Version]
  340. Alvarez-Erviti, L.; Rodriguez-Oroz, M.C.; Cooper, J.M.; Caballero, C.; Ferrer, I.; Obeso, J.A.; Schapira, A.H. Chaperone-mediated autophagy markers in Parkinson disease brains. Arch. Neurol. 2010, 67, 1464–1472. [Google Scholar] [CrossRef] [Green Version]
  341. Murphy, K.E.; Gysbers, A.M.; Abbott, S.K.; Spiro, A.S.; Furuta, A.; Cooper, A.; Garner, B.; Kabuta, T.; Halliday, G.M. Lysosomal-associated membrane protein 2 isoforms are differentially affected in early Parkinson’s disease. Mov. Disord. 2015, 30, 1639–1647. [Google Scholar] [CrossRef]
  342. Tanaka, Y.; Engelender, S.; Igarashi, S.; Rao, R.K.; Wanner, T.; Tanzi, R.E.; Sawa, A.; Valina, L.D.; Dawson, T.M.; Ross, C.A. Inducible expression of mutant alpha-synuclein decreases proteasome activity and increases sensitivity to mitochondria-dependent apoptosis. Hum. Mol. Genet. 2001, 10, 919–926. [Google Scholar] [CrossRef] [Green Version]
  343. Stefanis, L.; Larsen, K.E.; Rideout, H.J.; Sulzer, D.; Greene, L.A. Expression of A53T mutant but not wild-type alpha-synuclein in PC12 cells induces alterations of the ubiquitin-dependent degradation system, loss of dopamine release, and autophagic cell death. J. Neurosci. 2001, 21, 9549–9560. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  344. Martin-Clemente, B.; Alvarez-Castelao, B.; Mayo, I.; Sierra, A.B.; Diaz, V.; Milan, M.; Farinas, I.; Gomez-Isla, T.; Ferrer, I.; Castano, J.G. alpha-Synuclein expression levels do not significantly affect proteasome function and expression in mice and stably transfected PC12 cell lines. J. Biol. Chem. 2004, 279, 52984–52990. [Google Scholar] [CrossRef] [Green Version]
  345. Zondler, L.; Kostka, M.; Garidel, P.; Heinzelmann, U.; Hengerer, B.; Mayer, B.; Weishaupt, J.H.; Gillardon, F.; Danzer, K.M. Proteasome impairment by alpha-synuclein. PLoS ONE 2017, 12, e0184040. [Google Scholar] [CrossRef] [Green Version]
  346. Bentea, E.; Verbruggen, L.; Massie, A. The proteasome inhibition model of Parkinson’s disease. J. Parkinsons Dis. 2017, 7, 31–63. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  347. McNaught, K.S.; Jenner, P. Proteasomal function is impaired in substantia nigra in Parkinson’s disease. Neurosci. Lett. 2001, 297, 191–194. [Google Scholar] [CrossRef]
  348. McNaught, K.S.; Belizaire, R.; Isacson, O.; Jenner, P.; Olanow, C.W. Altered proteasomal function in sporadic Parkinson’s disease. Exp. Neurol. 2003, 179, 38–46. [Google Scholar] [CrossRef] [PubMed]
  349. Martinez-Vicente, M.; Sovak, G.; Cuervo, A.M. Protein degradation and aging. Exp. Gerontol. 2005, 40, 622–633. [Google Scholar] [CrossRef]
  350. Winslow, A.R.; Chen, C.W.; Corrochano, S.; Acevedo-Arozena, A.; Gordon, D.E.; Peden, A.A.; Lichtenberg, M.; Menzies, F.M.; Ravikumar, B.; Imarisio, S.; et al. Alpha-synuclein impairs macroautophagy: Implications for Parkinson’s disease. J. Cell Biol. 2010, 190, 1023–1037. [Google Scholar] [CrossRef] [Green Version]
  351. Zavodszky, E.; Seaman, M.N.; Moreau, K.; Jimenez-Sanchez, M.; Breusegem, S.Y.; Harbour, M.E.; Rubinsztein, D.C. Mutation in VPS35 associated with Parkinson’s disease impairs WASH complex association and inhibits autophagy. Nat. Commun. 2014, 5, 3828. [Google Scholar] [CrossRef] [PubMed]
  352. Yan, J.Q.; Yuan, Y.H.; Gao, Y.N.; Huang, J.Y.; Ma, K.L.; Gao, Y.; Zhang, W.Q.; Guo, X.F.; Chen, N.H. Overexpression of human E46K mutant alpha-synuclein impairs macroautophagy via inactivation of JNK1-Bcl-2 pathway. Mol. Neurobiol. 2014, 50, 685–701. [Google Scholar] [CrossRef] [PubMed]
  353. Lei, Z.; Cao, G.; Wei, G. A30P mutant alpha-synuclein impairs autophagic flux by inactivating JNK signaling to enhance ZKSCAN3 activity in midbrain dopaminergic neurons. Cell Death Dis 2019, 10, 133. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  354. Choubey, V.; Safiulina, D.; Vaarmann, A.; Cagalinec, M.; Wareski, P.; Kuum, M.; Zharkovsky, A.; Kaasik, A. Mutant A53T alpha-synuclein induces neuronal death by increasing mitochondrial autophagy. J. Biol. Chem. 2011, 286, 10814–10824. [Google Scholar] [CrossRef] [Green Version]
  355. Martinez-Vicente, M.; Talloczy, Z.; Kaushik, S.; Massey, A.C.; Mazzulli, J.; Mosharov, E.V.; Hodara, R.; Fredenburg, R.; Wu, D.C.; Follenzi, A.; et al. Dopamine-modified alpha-synuclein blocks chaperone-mediated autophagy. J. Clin. Investig. 2008, 118, 777–788. [Google Scholar] [CrossRef] [Green Version]
  356. Mazzulli, J.R.; Zunke, F.; Isacson, O.; Studer, L.; Krainc, D. alpha-Synuclein-induced lysosomal dysfunction occurs through disruptions in protein trafficking in human midbrain synucleinopathy models. Proc. Natl. Acad. Sci. USA 2016, 113, 1931–1936. [Google Scholar] [CrossRef] [Green Version]
  357. Osellame, L.D.; Rahim, A.A.; Hargreaves, I.P.; Gegg, M.E.; Richard-Londt, A.; Brandner, S.; Waddington, S.N.; Schapira, A.H.V.; Duchen, M.R. Mitochondria and quality control defects in a mouse model of Gaucher disease—Links to Parkinson’s disease. Cell Metab. 2013, 17, 941–953. [Google Scholar] [CrossRef] [Green Version]
  358. Sardi, S.P.; Clarke, J.; Viel, C.; Chan, M.; Tamsett, T.J.; Treleaven, C.M.; Bu, J.; Sweet, L.; Passini, M.A.; Dodge, J.C.; et al. Augmenting CNS glucocerebrosidase activity as a therapeutic strategy for parkinsonism and other Gaucher-related synucleinopathies. Proc. Natl. Acad. Sci. USA 2013, 110, 3537–3542. [Google Scholar] [CrossRef] [Green Version]
  359. Murphy, K.E.; Gysbers, A.M.; Abbott, S.K.; Tayebi, N.; Kim, W.S.; Sidransky, E.; Cooper, A.; Garner, B.; Halliday, G.M. Reduced glucocerebrosidase is associated with increased alpha-synuclein in sporadic Parkinson’s disease. Brain 2014, 137, 834–848. [Google Scholar] [CrossRef] [Green Version]
  360. Du, T.T.; Wang, L.; Duan, C.L.; Lu, L.L.; Zhang, J.L.; Gao, G.; Qiu, X.B.; Wang, X.M.; Yang, H. GBA deficiency promotes SNCA/alpha-synuclein accumulation through autophagic inhibition by inactivated PPP2A. Autophagy 2015, 11, 1803–1820. [Google Scholar] [CrossRef] [Green Version]
  361. Liu, G.; Chen, M.; Mi, N.; Yang, W.; Li, X.; Wang, P.; Yin, N.; Li, Y.; Yue, F.; Chan, P.; et al. Increased oligomerization and phosphorylation of alpha-synuclein are associated with decreased activity of glucocerebrosidase and protein phosphatase 2A in aging monkey brains. Neurobiol. Aging 2015, 36, 2649–2659. [Google Scholar] [CrossRef]
  362. Rocha, E.M.; Smith, G.A.; Park, E.; Cao, H.; Graham, A.R.; Brown, E.; McLean, J.R.; Hayes, M.A.; Beagan, J.; Izen, S.C.; et al. Sustained systemic glucocerebrosidase inhibition induces brain alpha-synuclein aggregation, microglia and complement C1q activation in mice. Antioxid. Redox. Signal. 2015, 23, 550–564. [Google Scholar] [CrossRef] [Green Version]
  363. Borghi, R.; Marchese, R.; Negro, A.; Marinelli, L.; Forloni, G.; Zaccheo, D.; Abbruzzese, G.; Tabaton, M. Full length alpha-synuclein is present in cerebrospinal fluid from Parkinson’s disease and normal subjects. Neurosci. Lett. 2000, 287, 65–67. [Google Scholar] [CrossRef]
  364. El-Agnaf, O.M.; Salem, S.A.; Paleologou, K.E.; Cooper, L.J.; Fullwood, N.J.; Gibson, M.J.; Curran, M.D.; Court, J.A.; Mann, D.M.; Ikeda, S.; et al. Alpha-synuclein implicated in Parkinson’s disease is present in extracellular biological fluids, including human plasma. FASEB J. 2003, 17, 1945–1947. [Google Scholar] [CrossRef]
  365. Lee, H.J.; Patel, S.; Lee, S.J. Intravesicular localization and exocytosis of alpha-synuclein and its aggregates. J. Neurosci. 2005, 25, 6016–6024. [Google Scholar] [CrossRef] [PubMed]
  366. Emmanouilidou, E.; Melachroinou, K.; Roumeliotis, T.; Garbis, S.D.; Ntzouni, M.; Margaritis, L.H.; Stefanis, L.; Vekrellis, K. Cell-produced alpha-synuclein is secreted in a calcium-dependent manner by exosomes and impacts neuronal survival. J. Neurosci. 2010, 30, 6838–6851. [Google Scholar] [CrossRef] [Green Version]
  367. Danzer, K.M.; Kranich, L.R.; Ruf, W.P.; Cagsal-Getkin, O.; Winslow, A.R.; Zhu, L.; Vanderburg, C.R.; McLean, P.J. Exosomal cell-to-cell transmission of alpha synuclein oligomers. Mol. Neurodegener. 2012, 7, 42. [Google Scholar] [CrossRef] [Green Version]
  368. Kordower, J.H.; Chu, Y.; Hauser, R.A.; Freeman, T.B.; Olanow, C.W. Lewy body-like pathology in long-term embryonic nigral transplants in Parkinson’s disease. Nat. Med. 2008, 14, 504–506. [Google Scholar] [CrossRef] [PubMed]
  369. Li, J.Y.; Englund, E.; Holton, J.L.; Soulet, D.; Hagell, P.; Lees, A.J.; Lashley, T.; Quinn, N.P.; Rehncrona, S.; Bjorklund, A.; et al. Lewy bodies in grafted neurons in subjects with Parkinson’s disease suggest host-to-graft disease propagation. Nat. Med. 2008, 14, 501–503. [Google Scholar] [CrossRef] [PubMed]
  370. Li, J.Y.; Englund, E.; Widner, H.; Rehncrona, S.; Bjorklund, A.; Lindvall, O.; Brundin, P. Characterization of Lewy body pathology in 12- and 16-year-old intrastriatal mesencephalic grafts surviving in a patient with Parkinson’s disease. Mov. Disord. 2010, 25, 1091–1096. [Google Scholar] [CrossRef]
  371. Kordower, J.H.; Dodiya, H.B.; Kordower, A.M.; Terpstra, B.; Paumier, K.; Madhavan, L.; Sortwell, C.; Steece-Collier, K.; Collier, T.J. Transfer of host-derived alpha synuclein to grafted dopaminergic neurons in rat. Neurobiol. Dis. 2011, 43, 552–557. [Google Scholar] [CrossRef] [Green Version]
  372. Kurowska, Z.; Englund, E.; Widner, H.; Lindvall, O.; Li, J.Y.; Brundin, P. Signs of degeneration in 12–22-year old grafts of mesencephalic dopamine neurons in patients with Parkinson’s disease. J. Parkinsons Dis. 2011, 1, 83–92. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  373. Masuda-Suzukake, M.; Nonaka, T.; Hosokawa, M.; Oikawa, T.; Arai, T.; Akiyama, H.; Mann, D.M.; Hasegawa, M. Prion-like spreading of pathological alpha-synuclein in brain. Brain 2013, 136, 1128–1138. [Google Scholar] [CrossRef]
  374. Recasens, A.; Dehay, B.; Bove, J.; Carballo-Carbajal, I.; Dovero, S.; Perez-Villalba, A.; Fernagut, P.O.; Blesa, J.; Parent, A.; Perier, C.; et al. Lewy body extracts from Parkinson disease brains trigger alpha-synuclein pathology and neurodegeneration in mice and monkeys. Ann. Neurol. 2014, 75, 351–362. [Google Scholar] [CrossRef]
  375. Luk, K.C.; Kehm, V.; Carroll, J.; Zhang, B.; O’Brien, P.; Trojanowski, J.Q.; Lee, V.M. Pathological alpha-synuclein transmission initiates Parkinson-like neurodegeneration in nontransgenic mice. Science 2012, 338, 949–953. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  376. Luk, K.C.; Song, C.; O’Brien, P.; Stieber, A.; Branch, J.R.; Brunden, K.R.; Trojanowski, J.Q.; Lee, V.M. Exogenous alpha-synuclein fibrils seed the formation of Lewy body-like intracellular inclusions in cultured cells. Proc. Natl. Acad. Sci. USA 2009, 106, 20051–20056. [Google Scholar] [CrossRef] [Green Version]
  377. Sacino, A.N.; Brooks, M.; Thomas, M.A.; McKinney, A.B.; Lee, S.; Regenhardt, R.W.; McGarvey, N.H.; Ayers, J.I.; Notterpek, L.; Borchelt, D.R.; et al. Intramuscular injection of alpha-synuclein induces CNS alpha-synuclein pathology and a rapid-onset motor phenotype in transgenic mice. Proc. Natl. Acad. Sci. USA 2014, 111, 10732–10737. [Google Scholar] [CrossRef] [Green Version]
  378. Paumier, K.L.; Luk, K.C.; Manfredsson, F.P.; Kanaan, N.M.; Lipton, J.W.; Collier, T.J.; Steece-Collier, K.; Kemp, C.J.; Celano, S.; Schulz, E.; et al. Intrastriatal injection of pre-formed mouse alpha-synuclein fibrils into rats triggers alpha-synuclein pathology and bilateral nigrostriatal degeneration. Neurobiol. Dis. 2015, 82, 185–199. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  379. Abdelmotilib, H.; Maltbie, T.; Delic, V.; Liu, Z.; Hu, X.; Fraser, K.B.; Moehle, M.S.; Stoyka, L.; Anabtawi, N.; Krendelchtchikova, V.; et al. alpha-Synuclein fibril-induced inclusion spread in rats and mice correlates with dopaminergic Neurodegeneration. Neurobiol. Dis. 2017, 105, 84–98. [Google Scholar] [CrossRef]
  380. Gustafsson, G.; Loov, C.; Persson, E.; Lazaro, D.F.; Takeda, S.; Bergstrom, J.; Erlandsson, A.; Sehlin, D.; Balaj, L.; Gyorgy, B.; et al. Secretion and uptake of alpha-synuclein via extracellular vesicles in cultured cells. Cell. Mol. Neurobiol. 2018, 38, 1539–1550. [Google Scholar] [CrossRef] [Green Version]
  381. Fussi, N.; Hollerhage, M.; Chakroun, T.; Nykanen, N.P.; Rosler, T.W.; Koeglsperger, T.; Wurst, W.; Behrends, C.; Hoglinger, G.U. Exosomal secretion of alpha-synuclein as protective mechanism after upstream blockage of macroautophagy. Cell Death Dis. 2018, 9, 757. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  382. El-Agnaf, O.M.; Salem, S.A.; Paleologou, K.E.; Curran, M.D.; Gibson, M.J.; Court, J.A.; Schlossmacher, M.G.; Allsop, D. Detection of oligomeric forms of alpha-synuclein protein in human plasma as a potential biomarker for Parkinson’s disease. FASEB J. 2006, 20, 419–425. [Google Scholar] [CrossRef]
  383. Delenclos, M.; Trendafilova, T.; Mahesh, D.; Baine, A.M.; Moussaud, S.; Yan, I.K.; Patel, T.; McLean, P.J. Investigation of endocytic pathways for the internalization of exosome-associated oligomeric alpha-synuclein. Front. Neurosci. 2017, 11, 172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  384. Sung, J.Y.; Kim, J.; Paik, S.R.; Park, J.H.; Ahn, Y.S.; Chung, K.C. Induction of neuronal cell death by Rab5A-dependent endocytosis of alpha-synuclein. J. Biol. Chem. 2001, 276, 27441–27448. [Google Scholar] [CrossRef] [Green Version]
  385. Lee, H.J.; Suk, J.E.; Bae, E.J.; Lee, J.H.; Paik, S.R.; Lee, S.J. Assembly-dependent endocytosis and clearance of extracellular alpha-synuclein. Int. J. Biochem. Cell Biol. 2008, 40, 1835–1849. [Google Scholar] [CrossRef] [PubMed]
  386. Ahn, K.J.; Paik, S.R.; Chung, K.C.; Kim, J. Amino acid sequence motifs and mechanistic features of the membrane translocation of alpha-synuclein. J. Neurochem. 2006, 97, 265–279. [Google Scholar] [CrossRef] [PubMed]
  387. Abounit, S.; Bousset, L.; Loria, F.; Zhu, S.; de Chaumont, F.; Pieri, L.; Olivo-Marin, J.C.; Melki, R.; Zurzolo, C. Tunneling nanotubes spread fibrillar alpha-synuclein by intercellular trafficking of lysosomes. EMBO J. 2016, 35, 2120–2138. [Google Scholar] [CrossRef]
  388. Kayed, R.; Sokolov, Y.; Edmonds, B.; McIntire, T.M.; Milton, S.C.; Hall, J.E.; Glabe, C.G. Permeabilization of lipid bilayers is a common conformation-dependent activity of soluble amyloid oligomers in protein misfolding diseases. J. Biol. Chem. 2004, 279, 46363–46366. [Google Scholar] [CrossRef] [Green Version]
  389. Jao, C.C.; Hegde, B.G.; Chen, J.; Haworth, I.S.; Langen, R. Structure of membrane-bound alpha-synuclein from site-directed spin labeling and computational refinement. Proc. Natl. Acad. Sci. USA 2008, 105, 19666–19671. [Google Scholar] [CrossRef] [Green Version]
  390. Mao, X.; Ou, M.T.; Karuppagounder, S.S.; Kam, T.I.; Yin, X.; Xiong, Y.; Ge, P.; Umanah, G.E.; Brahmachari, S.; Shin, J.H.; et al. Pathological alpha-synuclein transmission initiated by binding lymphocyte-activation gene 3. Science 2016, 353. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  391. Ihse, E.; Yamakado, H.; van Wijk, X.M.; Lawrence, R.; Esko, J.D.; Masliah, E. Cellular internalization of alpha-synuclein aggregates by cell surface heparan sulfate depends on aggregate conformation and cell type. Sci. Rep. 2017, 7, 9008. [Google Scholar] [CrossRef]
  392. Rey, N.L.; Steiner, J.A.; Maroof, N.; Luk, K.C.; Madaj, Z.; Trojanowski, J.Q.; Lee, V.M.; Brundin, P. Widespread transneuronal propagation of alpha-synucleinopathy triggered in olfactory bulb mimics prodromal Parkinson’s disease. J. Exp. Med. 2016, 213, 1759–1778. [Google Scholar] [CrossRef] [PubMed]
  393. Karampetsou, M.; Ardah, M.T.; Semitekolou, M.; Polissidis, A.; Samiotaki, M.; Kalomoiri, M.; Majbour, N.; Xanthou, G.; El-Agnaf, O.M.A.; Vekrellis, K. Phosphorylated exogenous alpha-synuclein fibrils exacerbate pathology and induce neuronal dysfunction in mice. Sci. Rep. 2017, 7, 16533. [Google Scholar] [CrossRef]
  394. Karpowicz, R.J., Jr.; Haney, C.M.; Mihaila, T.S.; Sandler, R.M.; Petersson, E.J.; Lee, V.M. Selective imaging of internalized proteopathic alpha-synuclein seeds in primary neurons reveals mechanistic insight into transmission of synucleinopathies. J. Biol. Chem. 2017, 292, 13482–13497. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  395. Luna, E.; Decker, S.C.; Riddle, D.M.; Caputo, A.; Zhang, B.; Cole, T.; Caswell, C.; Xie, S.X.; Lee, V.M.Y.; Luk, K.C. Differential alpha-synuclein expression contributes to selective vulnerability of hippocampal neuron subpopulations to fibril-induced toxicity. Acta Neuropathol. 2018, 135, 855–875. [Google Scholar] [CrossRef]
  396. Volpicelli-Daley, L.A.; Luk, K.C.; Lee, V.M. Addition of exogenous alpha-synuclein preformed fibrils to primary neuronal cultures to seed recruitment of endogenous alpha-synuclein to Lewy body and Lewy neurite-like aggregates. Nat. Protoc. 2014, 9, 2135–2146. [Google Scholar] [CrossRef] [Green Version]
  397. Sacino, A.N.; Brooks, M.; McKinney, A.B.; Thomas, M.A.; Shaw, G.; Golde, T.E.; Giasson, B.I. Brain injection of alpha-synuclein induces multiple proteinopathies, gliosis, and a neuronal injury marker. J. Neurosci. 2014, 34, 12368–12378. [Google Scholar] [CrossRef] [Green Version]
  398. Wu, Q.; Takano, H.; Riddle, D.M.; Trojanowski, J.Q.; Coulter, D.A.; Lee, V.M. Alpha-synuclein (alphaSyn) preformed fibrils induce endogenous alphaSyn aggregation, compromise synaptic activity and enhance synapse loss in cultured excitatory hippocampal neurons. J. Neurosci. 2019, 39, 5080–5094. [Google Scholar] [CrossRef] [Green Version]
  399. Chu, Y.; Kordower, J.H. The prion hypothesis of Parkinson’s disease. Curr. Neurol. Neurosci. Rep. 2015, 15, 28. [Google Scholar] [CrossRef] [PubMed]
  400. Uversky, V.N.; Li, J.; Fink, A.L. Evidence for a partially folded intermediate in alpha-synuclein fibril formation. J. Biol. Chem. 2001, 276, 10737–10744. [Google Scholar] [CrossRef] [Green Version]
  401. Wood, S.J.; Wypych, J.; Steavenson, S.; Louis, J.C.; Citron, M.; Biere, A.L. alpha-synuclein fibrillogenesis is nucleation-dependent. Implications for the pathogenesis of Parkinson’s disease. J. Biol. Chem. 1999, 274, 19509–19512. [Google Scholar] [CrossRef] [Green Version]
  402. Masuda-Suzukake, M.; Nonaka, T.; Hosokawa, M.; Kubo, M.; Shimozawa, A.; Akiyama, H.; Hasegawa, M. Pathological alpha-synuclein propagates through neural networks. Acta Neuropathol. Commun. 2014, 2, 88. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  403. Freundt, E.C.; Maynard, N.; Clancy, E.K.; Roy, S.; Bousset, L.; Sourigues, Y.; Covert, M.; Melki, R.; Kirkegaard, K.; Brahic, M. Neuron-to-neuron transmission of alpha-synuclein fibrils through axonal transport. Ann. Neurol. 2012, 72, 517–524. [Google Scholar] [CrossRef] [Green Version]
  404. Mezias, C.; Rey, N.; Brundin, P.; Raj, A. Neural connectivity predicts spreading of alpha-synuclein pathology in fibril-injected mouse models: Involvement of retrograde and anterograde axonal propagation. Neurobiol. Dis. 2020, 134, 104623. [Google Scholar] [CrossRef]
  405. Tang, Y.; Das, U.; Scott, D.A.; Roy, S. The slow axonal transport of alpha-synuclein--mechanistic commonalities amongst diverse cytosolic cargoes. Cytoskeleton 2012, 69, 506–513. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  406. Saha, A.R.; Hill, J.; Utton, M.A.; Asuni, A.A.; Ackerley, S.; Grierson, A.J.; Miller, C.C.; Davies, A.M.; Buchman, V.L.; Anderton, B.H.; et al. Parkinson’s disease alpha-synuclein mutations exhibit defective axonal transport in cultured neurons. J. Cell Sci. 2004, 117, 1017–1024. [Google Scholar] [CrossRef] [Green Version]
  407. Li, W.; Hoffman, P.N.; Stirling, W.; Price, D.L.; Lee, M.K. Axonal transport of human alpha-synuclein slows with aging but is not affected by familial Parkinson’s disease-linked mutations. J. Neurochem. 2004, 88, 401–410. [Google Scholar] [CrossRef] [PubMed]
  408. Prots, I.; Veber, V.; Brey, S.; Campioni, S.; Buder, K.; Riek, R.; Bohm, K.J.; Winner, B. alpha-Synuclein oligomers impair neuronal microtubule-kinesin interplay. J. Biol. Chem. 2013, 288, 21742–21754. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  409. Chu, Y.; Morfini, G.A.; Langhamer, L.B.; He, Y.; Brady, S.T.; Kordower, J.H. Alterations in axonal transport motor proteins in sporadic and experimental Parkinson’s disease. Brain 2012, 135, 2058–2073. [Google Scholar] [CrossRef] [Green Version]
  410. Chung, C.Y.; Koprich, J.B.; Siddiqi, H.; Isacson, O. Dynamic changes in presynaptic and axonal transport proteins combined with striatal neuroinflammation precede dopaminergic neuronal loss in a rat model of AAV alpha-synucleinopathy. J. Neurosci. 2009, 29, 3365–3373. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  411. Holec, S.A.M.; Woerman, A.L. Evidence of distinct alpha-synuclein strains underlying disease heterogeneity. Acta Neuropathol. 2020. [Google Scholar] [CrossRef] [PubMed]
  412. Lau, A.; So, R.W.L.; Lau, H.H.C.; Sang, J.C.; Ruiz-Riquelme, A.; Fleck, S.C.; Stuart, E.; Menon, S.; Visanji, N.P.; Meisl, G.; et al. alpha-Synuclein strains target distinct brain regions and cell types. Nat. Neurosci. 2020, 23, 21–31. [Google Scholar] [CrossRef]
  413. Peelaerts, W.; Bousset, L.; Van der Perren, A.; Moskalyuk, A.; Pulizzi, R.; Giugliano, M.; Van den Haute, C.; Melki, R.; Baekelandt, V. Alpha-synuclein strains cause distinct synucleinopathies after local and systemic administration. Nature 2015, 522, 340–344. [Google Scholar] [CrossRef]
  414. Rey, N.L.; Bousset, L.; George, S.; Madaj, Z.; Meyerdirk, L.; Schulz, E.; Steiner, J.A.; Melki, R.; Brundin, P. alpha-Synuclein conformational strains spread, seed and target neuronal cells differentially after injection into the olfactory bulb. Acta Neuropathol. Commun. 2019, 7, 221. [Google Scholar] [CrossRef] [Green Version]
  415. Wake, H.; Fields, R.D. Physiological function of microglia. Neuron Glia Biol. 2011, 7, 1–3. [Google Scholar] [CrossRef] [Green Version]
  416. Paolicelli, R.C.; Bolasco, G.; Pagani, F.; Maggi, L.; Scianni, M.; Panzanelli, P.; Giustetto, M.; Ferreira, T.A.; Guiducci, E.; Dumas, L.; et al. Synaptic pruning by microglia is necessary for normal brain development. Science 2011, 333, 1456–1458. [Google Scholar] [CrossRef] [Green Version]
  417. Sierra, A.; Encinas, J.M.; Deudero, J.J.; Chancey, J.H.; Enikolopov, G.; Overstreet-Wadiche, L.S.; Tsirka, S.E.; Maletic-Savatic, M. Microglia shape adult hippocampal neurogenesis through apoptosis-coupled phagocytosis. Cell Stem Cell 2010, 7, 483–495. [Google Scholar] [CrossRef] [Green Version]
  418. Nimmerjahn, A.; Kirchhoff, F.; Helmchen, F. Resting microglial cells are highly dynamic surveillants of brain parenchyma in vivo. Science 2005, 308, 1314–1318. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  419. Weinhard, L.; di Bartolomei, G.; Bolasco, G.; Machado, P.; Schieber, N.L.; Neniskyte, U.; Exiga, M.; Vadisiute, A.; Raggioli, A.; Schertel, A.; et al. Microglia remodel synapses by presynaptic trogocytosis and spine head filopodia induction. Nat. Commun. 2018, 9, 1228. [Google Scholar] [CrossRef] [Green Version]
  420. Colton, C.; Wilcock, D.M. Assessing activation states in microglia. CNS Neurol. Disord. Drug Targets 2010, 9, 174–191. [Google Scholar] [CrossRef]
  421. Kreutzberg, G.W. Microglia: A sensor for pathological events in the CNS. Trends Neurosci. 1996, 19, 312–318. [Google Scholar] [CrossRef]
  422. Streit, W.J.; Conde, J.R.; Fendrick, S.E.; Flanary, B.E.; Mariani, C.L. Role of microglia in the central nervous system’s immune response. Neurol. Res. 2005, 27, 685–691. [Google Scholar] [CrossRef]
  423. Van Rossum, D.; Hanisch, U.K. Microglia. Metab. Brain Dis. 2004, 19, 393–411. [Google Scholar] [CrossRef] [PubMed]
  424. Lecours, C.; Bordeleau, M.; Cantin, L.; Parent, M.; Paolo, T.D.; Tremblay, M.E. Microglial Implication in Parkinson’s Disease: Loss of Beneficial Physiological Roles or Gain of Inflammatory Functions? Front. Cell Neurosci. 2018, 12, 282. [Google Scholar] [CrossRef]
  425. Croisier, E.; Moran, L.B.; Dexter, D.T.; Pearce, R.K.; Graeber, M.B. Microglial inflammation in the parkinsonian substantia nigra: Relationship to alpha-synuclein deposition. J. Neuroinflamm. 2005, 2, 14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  426. Vieira, B.D.; Radford, R.A.; Chung, R.S.; Guillemin, G.J.; Pountney, D.L. Neuroinflammation in multiple system atrophy: Response to and cause of alpha-synuclein aggregation. Front. Cell Neurosci. 2015, 9, 437. [Google Scholar] [CrossRef]
  427. Surendranathan, A.; Su, L.; Mak, E.; Passamonti, L.; Hong, Y.T.; Arnold, R.; Vazquez Rodriguez, P.; Bevan-Jones, W.R.; Brain, S.A.E.; Fryer, T.D.; et al. Early microglial activation and peripheral inflammation in dementia with Lewy bodies. Brain 2018, 141, 3415–3427. [Google Scholar] [CrossRef] [PubMed]
  428. Mackenzie, I.R. Activated microglia in dementia with Lewy bodies. Neurology 2000, 55, 132–134. [Google Scholar] [CrossRef] [PubMed]
  429. Mosley, R.L.; Benner, E.J.; Kadiu, I.; Thomas, M.; Boska, M.D.; Hasan, K.; Laurie, C.; Gendelman, H.E. Neuroinflammation, oxidative stress and the pathogenesis of Parkinson’s disease. Clin. Neurosci. Res. 2006, 6, 261–281. [Google Scholar] [CrossRef] [Green Version]
  430. Appel, S.H.; Beers, D.R.; Henkel, J.S. T cell-microglial dialogue in Parkinson’s disease and amyotrophic lateral sclerosis: Are we listening? Trends Immunol. 2010, 31, 7–17. [Google Scholar] [CrossRef] [Green Version]
  431. Peterson, L.J.; Flood, P.M. Oxidative stress and microglial cells in Parkinson’s disease. Mediat. Inflamm. 2012, 2012, 401264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  432. Tang, Y.; Le, W. Differential roles of M1 and M2 microglia in neurodegenerative diseases. Mol. Neurobiol. 2016, 53, 1181–1194. [Google Scholar] [CrossRef]
  433. Liu, C.Y.; Wang, X.; Liu, C.; Zhang, H.L. Pharmacological targeting of microglial activation: New therapeutic approach. Front. Cell Neurosci. 2019, 13, 514. [Google Scholar] [CrossRef] [Green Version]
  434. Austin, S.A.; Floden, A.M.; Murphy, E.J.; Combs, C.K. Alpha-synuclein expression modulates microglial activation phenotype. J. Neurosci. 2006, 26, 10558–10563. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  435. Austin, S.A.; Rojanathammanee, L.; Golovko, M.Y.; Murphy, E.J.; Combs, C.K. Lack of alpha-synuclein modulates microglial phenotype in vitro. Neurochem. Res. 2011, 36, 994–1004. [Google Scholar] [CrossRef]
  436. Rojanathammanee, L.; Murphy, E.J.; Combs, C.K. Expression of mutant alpha-synuclein modulates microglial phenotype in vitro. J. Neuroinflamm. 2011, 8, 44. [Google Scholar] [CrossRef] [Green Version]
  437. Gardai, S.J.; Mao, W.; Schule, B.; Babcock, M.; Schoebel, S.; Lorenzana, C.; Alexander, J.; Kim, S.; Glick, H.; Hilton, K.; et al. Elevated alpha-synuclein impairs innate immune cell function and provides a potential peripheral biomarker for Parkinson’s disease. PLoS ONE 2013, 8, e71634. [Google Scholar] [CrossRef] [Green Version]
  438. Haenseler, W.; Zambon, F.; Lee, H.; Vowles, J.; Rinaldi, F.; Duggal, G.; Houlden, H.; Gwinn, K.; Wray, S.; Luk, K.C.; et al. Excess alpha-synuclein compromises phagocytosis in iPSC-derived macrophages. Sci. Rep. 2017, 7, 9003. [Google Scholar] [CrossRef]
  439. Kim, S.; Cho, S.H.; Kim, K.Y.; Shin, K.Y.; Kim, H.S.; Park, C.H.; Chang, K.A.; Lee, S.H.; Cho, D.; Suh, Y.H. Alpha-synuclein induces migration of BV-2 microglial cells by up-regulation of CD44 and MT1-MMP. J. Neurochem. 2009, 109, 1483–1496. [Google Scholar] [CrossRef]
  440. Streit, W.J.; Walter, S.A.; Pennell, N.A. Reactive microgliosis. Prog. Neurobiol. 1999, 57, 563–581. [Google Scholar] [CrossRef]
  441. Long-Smith, C.M.; Sullivan, A.M.; Nolan, Y.M. The influence of microglia on the pathogenesis of Parkinson’s disease. Prog. Neurobiol. 2009, 89, 277–287. [Google Scholar] [CrossRef] [PubMed]
  442. Halliday, G.M.; Stevens, C.H. Glia: Initiators and progressors of pathology in Parkinson’s disease. Mov. Disord. 2011, 26, 6–17. [Google Scholar] [CrossRef]
  443. De Araujo, F.M.; Cuenca-Bermejo, L.; Fernandez-Villalba, E.; Costa, S.L.; Silva, V.D.A.; Herrero, M.T. Role of microgliosis and NLRP3 inflammasome in Parkinson’s disease pathogenesis and therapy. Cell Mol. Neurobiol. 2021. [Google Scholar] [CrossRef]
  444. Jayaram, S.; Krishnamurthy, P.T. Role of microgliosis, oxidative stress and associated neuroinflammation in the pathogenesis of Parkinson’s disease: The therapeutic role of Nrf2 activators. Neurochem. Int. 2021, 145, 105014. [Google Scholar] [CrossRef]
  445. Marinova-Mutafchieva, L.; Sadeghian, M.; Broom, L.; Davis, J.B.; Medhurst, A.D.; Dexter, D.T. Relationship between microglial activation and dopaminergic neuronal loss in the substantia nigra: A time course study in a 6-hydroxydopamine model of Parkinson’s disease. J. Neurochem. 2009, 110, 966–975. [Google Scholar] [CrossRef]
  446. Ouchi, Y.; Yoshikawa, E.; Sekine, Y.; Futatsubashi, M.; Kanno, T.; Ogusu, T.; Torizuka, T. Microglial activation and dopamine terminal loss in early Parkinson’s disease. Ann. Neurol. 2005, 57, 168–175. [Google Scholar] [CrossRef]
  447. Bartels, A.L.; Willemsen, A.T.; Doorduin, J.; de Vries, E.F.; Dierckx, R.A.; Leenders, K.L. [11C]-PK11195 PET: Quantification of neuroinflammation and a monitor of anti-inflammatory treatment in Parkinson’s disease? Parkinsonism Relat. Disord. 2010, 16, 57–59. [Google Scholar] [CrossRef] [PubMed]
  448. Gerhard, A.; Pavese, N.; Hotton, G.; Turkheimer, F.; Es, M.; Hammers, A.; Eggert, K.; Oertel, W.; Banati, R.B.; Brooks, D.J. In vivo imaging of microglial activation with [11C](R)-PK11195 PET in idiopathic Parkinson’s disease. Neurobiol. Dis. 2006, 21, 404–412. [Google Scholar] [CrossRef]
  449. Barkholt, P.; Sanchez-Guajardo, V.; Kirik, D.; Romero-Ramos, M. Long-term polarization of microglia upon alpha-synuclein overexpression in nonhuman primates. Neuroscience 2012, 208, 85–96. [Google Scholar] [CrossRef]
  450. Sanchez-Guajardo, V.; Febbraro, F.; Kirik, D.; Romero-Ramos, M. Microglia acquire distinct activation profiles depending on the degree of alpha-synuclein neuropathology in a rAAV based model of Parkinson’s disease. PLoS ONE 2010, 5, e8784. [Google Scholar] [CrossRef] [PubMed]
  451. Ginns, E.I.; Mak, S.K.; Ko, N.; Karlgren, J.; Akbarian, S.; Chou, V.P.; Guo, Y.; Lim, A.; Samuelsson, S.; LaMarca, M.L.; et al. Neuroinflammation and alpha-synuclein accumulation in response to glucocerebrosidase deficiency are accompanied by synaptic dysfunction. Mol. Genet. Metab. 2014, 111, 152–162. [Google Scholar] [CrossRef]
  452. Gomez-Isla, T.; Irizarry, M.C.; Mariash, A.; Cheung, B.; Soto, O.; Schrump, S.; Sondel, J.; Kotilinek, L.; Day, J.; Schwarzschild, M.A.; et al. Motor dysfunction and gliosis with preserved dopaminergic markers in human alpha-synuclein A30P transgenic mice. Neurobiol. Aging 2003, 24, 245–258. [Google Scholar] [CrossRef]
  453. Emmer, K.L.; Waxman, E.A.; Covy, J.P.; Giasson, B.I. E46K human alpha-synuclein transgenic mice develop Lewy-like and tau pathology associated with age-dependent, detrimental motor impairment. J. Biol. Chem. 2011, 286, 35104–35118. [Google Scholar] [CrossRef] [Green Version]
  454. Watson, M.B.; Richter, F.; Lee, S.K.; Gabby, L.; Wu, J.; Masliah, E.; Effros, R.B.; Chesselet, M.F. Regionally-specific microglial activation in young mice over-expressing human wildtype alpha-synuclein. Exp. Neurol. 2012, 237, 318–334. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  455. Garcia-Esparcia, P.; Llorens, F.; Carmona, M.; Ferrer, I. Complex deregulation and expression of cytokines and mediators of the immune response in Parkinson’s disease brain is region dependent. Brain Pathol. 2014, 24, 584–598. [Google Scholar] [CrossRef]
  456. Zhang, W.; Wang, T.; Pei, Z.; Miller, D.S.; Wu, X.; Block, M.L.; Wilson, B.; Zhou, Y.; Hong, J.S.; Zhang, J. Aggregated alpha-synuclein activates microglia: A process leading to disease progression in Parkinson’s disease. FASEB J. 2005, 19, 533–542. [Google Scholar] [CrossRef] [PubMed]
  457. Ishizawa, K.; Komori, T.; Sasaki, S.; Arai, N.; Mizutani, T.; Hirose, T. Microglial activation parallels system degeneration in multiple system atrophy. J. Neuropathol. Exp. Neurol. 2004, 63, 43–52. [Google Scholar] [CrossRef] [Green Version]
  458. Kubler, D.; Wachter, T.; Cabanel, N.; Su, Z.; Turkheimer, F.E.; Dodel, R.; Brooks, D.J.; Oertel, W.H.; Gerhard, A. Widespread microglial activation in multiple system atrophy. Mov. Disord. 2019, 34, 564–568. [Google Scholar] [CrossRef] [Green Version]
  459. Harms, A.S.; Kordower, J.H.; Sette, A.; Lindestam Arlehamn, C.S.; Sulzer, D.; Mach, R.H. Inflammation in experimental models of alpha-synucleinopathies. Mov. Disord. 2021, 36, 37–49. [Google Scholar] [CrossRef]
  460. Hirsch, E.C.; Hunot, S.; Damier, P.; Faucheux, B. Glial cells and inflammation in Parkinson’s disease: A role in neurodegeneration? Ann. Neurol. 1998, 44, S115–S120. [Google Scholar] [CrossRef]
  461. Shavali, S.; Combs, C.K.; Ebadi, M. Reactive macrophages increase oxidative stress and alpha-synuclein nitration during death of dopaminergic neuronal cells in co-culture: Relevance to Parkinson’s disease. Neurochem. Res. 2006, 31, 85–94. [Google Scholar] [CrossRef] [PubMed]
  462. Gao, H.M.; Kotzbauer, P.T.; Uryu, K.; Leight, S.; Trojanowski, J.Q.; Lee, V.M. Neuroinflammation and oxidation/nitration of alpha-synuclein linked to dopaminergic neurodegeneration. J. Neurosci. 2008, 28, 7687–7698. [Google Scholar] [CrossRef] [PubMed]
  463. Zhang, W.; Dallas, S.; Zhang, D.; Guo, J.P.; Pang, H.; Wilson, B.; Miller, D.S.; Chen, B.; McGeer, P.L.; Hong, J.S.; et al. Microglial PHOX and Mac-1 are essential to the enhanced dopaminergic neurodegeneration elicited by A30P and A53T mutant alpha-synuclein. Glia 2007, 55, 1178–1188. [Google Scholar] [CrossRef]
  464. Bick, R.J.; Poindexter, B.J.; Kott, M.M.; Liang, Y.A.; Dinh, K.; Kaur, B.; Bick, D.L.; Doursout, M.F.; Schiess, M.C. Cytokines disrupt intracellular patterns of Parkinson’s disease-associated proteins alpha-synuclein, tau and ubiquitin in cultured glial cells. Brain Res. 2008, 1217, 203–212. [Google Scholar] [CrossRef]
  465. Schiess, M.C.; Barnes, J.L.; Ellmore, T.M.; Poindexter, B.J.; Dinh, K.; Bick, R.J. CSF from Parkinson disease patients differentially affects cultured microglia and astrocytes. BMC Neurosci. 2010, 11, 151. [Google Scholar] [CrossRef] [Green Version]
  466. Lee, M.K.; Stirling, W.; Xu, Y.; Xu, X.; Qui, D.; Mandir, A.S.; Dawson, T.M.; Copeland, N.G.; Jenkins, N.A.; Price, D.L. Human alpha-synuclein-harboring familial Parkinson’s disease-linked Ala-53 --> Thr mutation causes neurodegenerative disease with alpha-synuclein aggregation in transgenic mice. Proc. Natl. Acad. Sci. USA 2002, 99, 8968–8973. [Google Scholar] [CrossRef] [Green Version]
  467. Miller, R.M.; Kiser, G.L.; Kaysser-Kranich, T.; Casaceli, C.; Colla, E.; Lee, M.K.; Palaniappan, C.; Federoff, H.J. Wild-type and mutant alpha-synuclein induce a multi-component gene expression profile consistent with shared pathophysiology in different transgenic mouse models of PD. Exp. Neurol. 2007, 204, 421–432. [Google Scholar] [CrossRef]
  468. Couch, Y.; Alvarez-Erviti, L.; Sibson, N.R.; Wood, M.J.; Anthony, D.C. The acute inflammatory response to intranigral alpha-synuclein differs significantly from intranigral lipopolysaccharide and is exacerbated by peripheral inflammation. J. Neuroinflamm. 2011, 8, 166. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  469. Wilms, H.; Rosenstiel, P.; Romero-Ramos, M.; Arlt, A.; Schafer, H.; Seegert, D.; Kahle, P.J.; Odoy, S.; Claasen, J.H.; Holzknecht, C.; et al. Suppression of MAP kinases inhibits microglial activation and attenuates neuronal cell death induced by alpha-synuclein protofibrils. Int. J. Immunopathol. Pharmacol. 2009, 22, 897–909. [Google Scholar] [CrossRef]
  470. Jin, J.; Shie, F.S.; Liu, J.; Wang, Y.; Davis, J.; Schantz, A.M.; Montine, K.S.; Montine, T.J.; Zhang, J. Prostaglandin E2 receptor subtype 2 (EP2) regulates microglial activation and associated neurotoxicity induced by aggregated alpha-synuclein. J. Neuroinflamm. 2007, 4, 2. [Google Scholar] [CrossRef]
  471. Theodore, S.; Cao, S.; McLean, P.J.; Standaert, D.G. Targeted overexpression of human alpha-synuclein triggers microglial activation and an adaptive immune response in a mouse model of Parkinson disease. J. Neuropathol. Exp. Neurol. 2008, 67, 1149–1158. [Google Scholar] [CrossRef] [Green Version]
  472. St Martin, J.L.; Klucken, J.; Outeiro, T.F.; Nguyen, P.; Keller-McGandy, C.; Cantuti-Castelvetri, I.; Grammatopoulos, T.N.; Standaert, D.G.; Hyman, B.T.; McLean, P.J. Dopaminergic neuron loss and up-regulation of chaperone protein mRNA induced by targeted over-expression of alpha-synuclein in mouse substantia nigra. J. Neurochem. 2007, 100, 1449–1457. [Google Scholar] [CrossRef]
  473. Alvarez-Erviti, L.; Couch, Y.; Richardson, J.; Cooper, J.M.; Wood, M.J. Alpha-synuclein release by neurons activates the inflammatory response in a microglial cell line. Neurosci. Res. 2011, 69, 337–342. [Google Scholar] [CrossRef] [PubMed]
  474. Park, J.Y.; Paik, S.R.; Jou, I.; Park, S.M. Microglial phagocytosis is enhanced by monomeric alpha-synuclein, not aggregated alpha-synuclein: Implications for Parkinson’s disease. Glia 2008, 56, 1215–1223. [Google Scholar] [CrossRef] [PubMed]
  475. Lee, S.B.; Park, S.M.; Ahn, K.J.; Chung, K.C.; Paik, S.R.; Kim, J. Identification of the amino acid sequence motif of alpha-synuclein responsible for macrophage activation. Biochem. Biophys. Res. Commun. 2009, 381, 39–43. [Google Scholar] [CrossRef]
  476. Klegeris, A.; Pelech, S.; Giasson, B.I.; Maguire, J.; Zhang, H.; McGeer, E.G.; McGeer, P.L. Alpha-synuclein activates stress signaling protein kinases in THP-1 cells and microglia. Neurobiol. Aging 2008, 29, 739–752. [Google Scholar] [CrossRef] [PubMed]
  477. Su, X.; Maguire-Zeiss, K.A.; Giuliano, R.; Prifti, L.; Venkatesh, K.; Federoff, H.J. Synuclein activates microglia in a model of Parkinson’s disease. Neurobiol. Aging 2008, 29, 1690–1701. [Google Scholar] [CrossRef] [Green Version]
  478. Lee, E.J.; Woo, M.S.; Moon, P.G.; Baek, M.C.; Choi, I.Y.; Kim, W.K.; Junn, E.; Kim, H.S. Alpha-synuclein activates microglia by inducing the expressions of matrix metalloproteinases and the subsequent activation of protease-activated receptor-1. J. Immunol. 2010, 185, 615–623. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  479. Roodveldt, C.; Labrador-Garrido, A.; Gonzalez-Rey, E.; Fernandez-Montesinos, R.; Caro, M.; Lachaud, C.C.; Waudby, C.A.; Delgado, M.; Dobson, C.M.; Pozo, D. Glial innate immunity generated by non-aggregated alpha-synuclein in mouse: Differences between wild-type and Parkinson’s disease-linked mutants. PLoS ONE 2010, 5, e13481. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  480. Li, N.; Stewart, T.; Sheng, L.; Shi, M.; Cilento, E.M.; Wu, Y.; Hong, J.S.; Zhang, J. Immunoregulation of microglial polarization: An unrecognized physiological function of alpha-synuclein. J. Neuroinflamm. 2020, 17, 272. [Google Scholar] [CrossRef]
  481. Reynolds, A.D.; Glanzer, J.G.; Kadiu, I.; Ricardo-Dukelow, M.; Chaudhuri, A.; Ciborowski, P.; Cerny, R.; Gelman, B.; Thomas, M.P.; Mosley, R.L.; et al. Nitrated alpha-synuclein-activated microglial profiling for Parkinson’s disease. J. Neurochem. 2008, 104, 1504–1525. [Google Scholar] [CrossRef] [Green Version]
  482. Beraud, D.; Twomey, M.; Bloom, B.; Mittereder, A.; Ton, V.; Neitzke, K.; Chasovskikh, S.; Mhyre, T.R.; Maguire-Zeiss, K.A. Alpha-synuclein alters toll-like receptor expression. Front. Neurosci. 2011, 5, 80. [Google Scholar] [CrossRef] [Green Version]
  483. Thomas, M.P.; Chartrand, K.; Reynolds, A.; Vitvitsky, V.; Banerjee, R.; Gendelman, H.E. Ion channel blockade attenuates aggregated alpha synuclein induction of microglial reactive oxygen species: Relevance for the pathogenesis of Parkinson’s disease. J. Neurochem. 2007, 100, 503–519. [Google Scholar] [CrossRef]
  484. Hoffmann, A.; Ettle, B.; Bruno, A.; Kulinich, A.; Hoffmann, A.C.; von Wittgenstein, J.; Winkler, J.; Xiang, W.; Schlachetzki, J.C.M. Alpha-synuclein activates BV2 microglia dependent on its aggregation state. Biochem. Biophys. Res. Commun. 2016, 479, 881–886. [Google Scholar] [CrossRef]
  485. Hoenen, C.; Gustin, A.; Birck, C.; Kirchmeyer, M.; Beaume, N.; Felten, P.; Grandbarbe, L.; Heuschling, P.; Heurtaux, T. Alpha-synuclein proteins promote pro-inflammatory cascades in microglia: Stronger effects of the A53T mutant. PLoS ONE 2016, 11, e0162717. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  486. Li, Y.; Niu, M.; Zhao, A.; Kang, W.; Chen, Z.; Luo, N.; Zhou, L.; Zhu, X.; Lu, L.; Liu, J. CXCL12 is involved in alpha-synuclein-triggered neuroinflammation of Parkinson’s disease. J. Neuroinflamm. 2019, 16, 263. [Google Scholar] [CrossRef] [PubMed]
  487. Liu, J.; Zhou, Y.; Wang, Y.; Fong, H.; Murray, T.M.; Zhang, J. Identification of proteins involved in microglial endocytosis of alpha-synuclein. J. Proteome Res. 2007, 6, 3614–3627. [Google Scholar] [CrossRef]
  488. Hou, L.; Bao, X.; Zang, C.; Yang, H.; Sun, F.; Che, Y.; Wu, X.; Li, S.; Zhang, D.; Wang, Q. Integrin CD11b mediates alpha-synuclein-induced activation of NADPH oxidase through a Rho-dependent pathway. Redox. Biol. 2018, 14, 600–608. [Google Scholar] [CrossRef]
  489. Hou, L.; Qu, X.; Qiu, X.; Huang, R.; Zhao, X.; Wang, Q. Integrin CD11b mediates locus coeruleus noradrenergic neurodegeneration in a mouse Parkinson’s disease model. J. Neuroinflamm. 2020, 17, 148. [Google Scholar] [CrossRef] [PubMed]
  490. Wang, S.; Chu, C.H.; Stewart, T.; Ginghina, C.; Wang, Y.; Nie, H.; Guo, M.; Wilson, B.; Hong, J.S.; Zhang, J. alpha-Synuclein, a chemoattractant, directs microglial migration via H2O2-dependent Lyn phosphorylation. Proc. Natl. Acad. Sci. USA 2015, 112, E1926–E1935. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  491. Su, X.; Federoff, H.J.; Maguire-Zeiss, K.A. Mutant alpha-synuclein overexpression mediates early proinflammatory activity. Neurotox. Res. 2009, 16, 238–254. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  492. Jiang, T.; Hoekstra, J.; Heng, X.; Kang, W.; Ding, J.; Liu, J.; Chen, S.; Zhang, J. P2X7 receptor is critical in alpha-synuclein—Mediated microglial NADPH oxidase activation. Neurobiol. Aging 2015, 36, 2304–2318. [Google Scholar] [CrossRef]
  493. Cao, S.; Standaert, D.G.; Harms, A.S. The gamma chain subunit of Fc receptors is required for alpha-synuclein-induced pro-inflammatory signaling in microglia. J. Neuroinflamm. 2012, 9, 259. [Google Scholar] [CrossRef] [Green Version]
  494. Stefanova, N.; Fellner, L.; Reindl, M.; Masliah, E.; Poewe, W.; Wenning, G.K. Toll-like receptor 4 promotes alpha-synuclein clearance and survival of nigral dopaminergic neurons. Am. J. Pathol. 2011, 179, 954–963. [Google Scholar] [CrossRef] [Green Version]
  495. Venezia, S.; Refolo, V.; Polissidis, A.; Stefanis, L.; Wenning, G.K.; Stefanova, N. Toll-like receptor 4 stimulation with monophosphoryl lipid A ameliorates motor deficits and nigral neurodegeneration triggered by extraneuronal alpha-synucleinopathy. Mol. Neurodegener. 2017, 12, 52. [Google Scholar] [CrossRef] [PubMed]
  496. Fellner, L.; Irschick, R.; Schanda, K.; Reindl, M.; Klimaschewski, L.; Poewe, W.; Wenning, G.K.; Stefanova, N. Toll-like receptor 4 is required for alpha-synuclein dependent activation of microglia and astroglia. Glia 2013, 61, 349–360. [Google Scholar] [CrossRef] [Green Version]
  497. Kim, C.; Ho, D.H.; Suk, J.E.; You, S.; Michael, S.; Kang, J.; Joong Lee, S.; Masliah, E.; Hwang, D.; Lee, H.J.; et al. Neuron-released oligomeric alpha-synuclein is an endogenous agonist of TLR2 for paracrine activation of microglia. Nat. Commun. 2013, 4, 1562. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  498. Dzamko, N.; Gysbers, A.; Perera, G.; Bahar, A.; Shankar, A.; Gao, J.; Fu, Y.; Halliday, G.M. Toll-like receptor 2 is increased in neurons in Parkinson’s disease brain and may contribute to alpha-synuclein pathology. Acta Neuropathol. 2017, 133, 303–319. [Google Scholar] [CrossRef] [Green Version]
  499. Kim, C.; Spencer, B.; Rockenstein, E.; Yamakado, H.; Mante, M.; Adame, A.; Fields, J.A.; Masliah, D.; Iba, M.; Lee, H.J.; et al. Immunotherapy targeting toll-like receptor 2 alleviates neurodegeneration in models of synucleinopathy by modulating alpha-synuclein transmission and neuroinflammation. Mol. Neurodegener. 2018, 13, 43. [Google Scholar] [CrossRef]
  500. Kwon, S.; Iba, M.; Masliah, E.; Kim, C. Targeting microglial and neuronal toll-like receptor 2 in synucleinopathies. Exp. Neurobiol. 2019, 28, 547–553. [Google Scholar] [CrossRef] [Green Version]
  501. Caplan, I.F.; Maguire-Zeiss, K.A. Toll-like receptor 2 signaling and current approaches for therapeutic modulation in synucleinopathies. Front. Pharmacol. 2018, 9, 417. [Google Scholar] [CrossRef] [Green Version]
  502. Roodveldt, C.; Labrador-Garrido, A.; Gonzalez-Rey, E.; Lachaud, C.C.; Guilliams, T.; Fernandez-Montesinos, R.; Benitez-Rondan, A.; Robledo, G.; Hmadcha, A.; Delgado, M.; et al. Preconditioning of microglia by alpha-synuclein strongly affects the response induced by toll-like receptor (TLR) stimulation. PLoS ONE 2013, 8, e79160. [Google Scholar] [CrossRef] [Green Version]
  503. Reynolds, A.D.; Kadiu, I.; Garg, S.K.; Glanzer, J.G.; Nordgren, T.; Ciborowski, P.; Banerjee, R.; Gendelman, H.E. Nitrated alpha-synuclein and microglial neuroregulatory activities. J. Neuroimmune Pharmacol. 2008, 3, 59–74. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  504. Daniele, S.G.; Beraud, D.; Davenport, C.; Cheng, K.; Yin, H.; Maguire-Zeiss, K.A. Activation of MyD88-dependent TLR1/2 signaling by misfolded alpha-synuclein, a protein linked to neurodegenerative disorders. Sci. Signal. 2015, 8, ra45. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  505. Prabhakaran, K.; Chapman, G.D.; Gunasekar, P.G. alpha-Synuclein overexpression enhances manganese-induced neurotoxicity through the NF-kappaB-mediated pathway. Toxicol. Mech. Methods 2011, 21, 435–443. [Google Scholar] [CrossRef]
  506. Maekawa, T.; Sasaoka, T.; Azuma, S.; Ichikawa, T.; Melrose, H.L.; Farrer, M.J.; Obata, F. Leucine-rich repeat kinase 2 (LRRK2) regulates alpha-synuclein clearance in microglia. BMC Neurosci. 2016, 17, 77. [Google Scholar] [CrossRef] [Green Version]
  507. Daher, J.P.; Volpicelli-Daley, L.A.; Blackburn, J.P.; Moehle, M.S.; West, A.B. Abrogation of alpha-synuclein-mediated dopaminergic neurodegeneration in LRRK2-deficient rats. Proc. Natl. Acad. Sci. USA 2014, 111, 9289–9294. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  508. Codolo, G.; Plotegher, N.; Pozzobon, T.; Brucale, M.; Tessari, I.; Bubacco, L.; de Bernard, M. Triggering of inflammasome by aggregated alpha-synuclein, an inflammatory response in synucleinopathies. PLoS ONE 2013, 8, e55375. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  509. Freeman, D.; Cedillos, R.; Choyke, S.; Lukic, Z.; McGuire, K.; Marvin, S.; Burrage, A.M.; Sudholt, S.; Rana, A.; O’Connor, C.; et al. Alpha-synuclein induces lysosomal rupture and cathepsin dependent reactive oxygen species following endocytosis. PLoS ONE 2013, 8, e62143. [Google Scholar] [CrossRef] [PubMed]
  510. Zhou, Y.; Lu, M.; Du, R.H.; Qiao, C.; Jiang, C.Y.; Zhang, K.Z.; Ding, J.H.; Hu, G. MicroRNA-7 targets Nod-like receptor protein 3 inflammasome to modulate neuroinflammation in the pathogenesis of Parkinson’s disease. Mol. Neurodegener. 2016, 11, 28. [Google Scholar] [CrossRef] [Green Version]
  511. Beraud, D.; Hathaway, H.A.; Trecki, J.; Chasovskikh, S.; Johnson, D.A.; Johnson, J.A.; Federoff, H.J.; Shimoji, M.; Mhyre, T.R.; Maguire-Zeiss, K.A. Microglial activation and antioxidant responses induced by the Parkinson’s disease protein alpha-synuclein. J. Neuroimmune Pharmacol. 2013, 8, 94–117. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  512. Lastres-Becker, I.; Ulusoy, A.; Innamorato, N.G.; Sahin, G.; Rabano, A.; Kirik, D.; Cuadrado, A. alpha-Synuclein expression and Nrf2 deficiency cooperate to aggravate protein aggregation, neuronal death and inflammation in early-stage Parkinson’s disease. Hum. Mol. Genet. 2012, 21, 3173–3192. [Google Scholar] [CrossRef] [Green Version]
  513. Gan, L.; Vargas, M.R.; Johnson, D.A.; Johnson, J.A. Astrocyte-specific overexpression of Nrf2 delays motor pathology and synuclein aggregation throughout the CNS in the alpha-synuclein mutant (A53T) mouse model. J. Neurosci. 2012, 32, 17775–17787. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  514. Skibinski, G.; Hwang, V.; Ando, D.M.; Daub, A.; Lee, A.K.; Ravisankar, A.; Modan, S.; Finucane, M.M.; Shaby, B.A.; Finkbeiner, S. Nrf2 mitigates LRRK2- and alpha-synuclein-induced neurodegeneration by modulating proteostasis. Proc. Natl. Acad. Sci. USA 2017, 114, 1165–1170. [Google Scholar] [CrossRef] [Green Version]
  515. Subbarayan, M.S.; Hudson, C.; Moss, L.D.; Nash, K.R.; Bickford, P.C. T cell infiltration and upregulation of MHCII in microglia leads to accelerated neuronal loss in an alpha-synuclein rat model of Parkinson’s disease. J. Neuroinflamm. 2020, 17, 242. [Google Scholar] [CrossRef]
  516. Harms, A.S.; Cao, S.; Rowse, A.L.; Thome, A.D.; Li, X.; Mangieri, L.R.; Cron, R.Q.; Shacka, J.J.; Raman, C.; Standaert, D.G. MHCII is required for alpha-synuclein-induced activation of microglia, CD4 T cell proliferation, and dopaminergic neurodegeneration. J. Neurosci. 2013, 33, 9592–9600. [Google Scholar] [CrossRef]
  517. Williams, G.P.; Marmion, D.J.; Schonhoff, A.M.; Jurkuvenaite, A.; Won, W.J.; Standaert, D.G.; Kordower, J.H.; Harms, A.S. T cell infiltration in both human multiple system atrophy and a novel mouse model of the disease. Acta Neuropathol. 2020, 139, 855–874. [Google Scholar] [CrossRef] [Green Version]
  518. Lindestam Arlehamn, C.S.; Dhanwani, R.; Pham, J.; Kuan, R.; Frazier, A.; Rezende Dutra, J.; Phillips, E.; Mallal, S.; Roederer, M.; Marder, K.S.; et al. Alpha-synuclein-specific T cell reactivity is associated with preclinical and early Parkinson’s disease. Nat. Commun. 2020, 11, 1875. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  519. Sulzer, D.; Alcalay, R.N.; Garretti, F.; Cote, L.; Kanter, E.; Agin-Liebes, J.; Liong, C.; McMurtrey, C.; Hildebrand, W.H.; Mao, X.; et al. T cells from patients with Parkinson’s disease recognize alpha-synuclein peptides. Nature 2017, 546, 656–661. [Google Scholar] [CrossRef] [Green Version]
  520. Jimenez-Ferrer, I.; Backstrom, F.; Duenas-Rey, A.; Jewett, M.; Boza-Serrano, A.; Luk, K.C.; Deierborg, T.; Swanberg, M. The MHC class II transactivator modulates seeded alpha-synuclein pathology and dopaminergic neurodegeneration in an in vivo rat model of Parkinson’s disease. Brain Behav. Immun. 2021, 91, 369–382. [Google Scholar] [CrossRef]
  521. Choi, I.; Zhang, Y.; Seegobin, S.P.; Pruvost, M.; Wang, Q.; Purtell, K.; Zhang, B.; Yue, Z. Microglia clear neuron-released alpha-synuclein via selective autophagy and prevent neurodegeneration. Nat. Commun. 2020, 11, 1386. [Google Scholar] [CrossRef] [Green Version]
  522. Lee, H.J.; Suk, J.E.; Bae, E.J.; Lee, S.J. Clearance and deposition of extracellular alpha-synuclein aggregates in microglia. Biochem. Biophys. Res. Commun. 2008, 372, 423–428. [Google Scholar] [CrossRef]
  523. Nash, Y.; Schmukler, E.; Trudler, D.; Pinkas-Kramarski, R.; Frenkel, D. DJ-1 deficiency impairs autophagy and reduces alpha-synuclein phagocytosis by microglia. J. Neurochem. 2017, 143, 584–594. [Google Scholar] [CrossRef] [Green Version]
  524. Wu, K.C.; Liou, H.H.; Kao, Y.H.; Lee, C.Y.; Lin, C.J. The critical role of Nramp1 in degrading alpha-synuclein oligomers in microglia under iron overload condition. Neurobiol. Dis. 2017, 104, 61–72. [Google Scholar] [CrossRef] [PubMed]
  525. Bussi, C.; Peralta Ramos, J.M.; Arroyo, D.S.; Gallea, J.I.; Ronchi, P.; Kolovou, A.; Wang, J.M.; Florey, O.; Celej, M.S.; Schwab, Y.; et al. Alpha-synuclein fibrils recruit TBK1 and OPTN to lysosomal damage sites and induce autophagy in microglial cells. J. Cell Sci. 2018, 131. [Google Scholar] [CrossRef] [Green Version]
  526. Bliederhaeuser, C.; Grozdanov, V.; Speidel, A.; Zondler, L.; Ruf, W.P.; Bayer, H.; Kiechle, M.; Feiler, M.S.; Freischmidt, A.; Brenner, D.; et al. Age-dependent defects of alpha-synuclein oligomer uptake in microglia and monocytes. Acta Neuropathol. 2016, 131, 379–391. [Google Scholar] [CrossRef]
  527. Scheffold, A.; Holtman, I.R.; Dieni, S.; Brouwer, N.; Katz, S.F.; Jebaraj, B.M.; Kahle, P.J.; Hengerer, B.; Lechel, A.; Stilgenbauer, S.; et al. Telomere shortening leads to an acceleration of synucleinopathy and impaired microglia response in a genetic mouse model. Acta Neuropathol. Commun. 2016, 4, 87. [Google Scholar] [CrossRef] [Green Version]
  528. Park, J.Y.; Kim, K.S.; Lee, S.B.; Ryu, J.S.; Chung, K.C.; Choo, Y.K.; Jou, I.; Kim, J.; Park, S.M. On the mechanism of internalization of alpha-synuclein into microglia: Roles of ganglioside GM1 and lipid raft. J. Neurochem. 2009, 110, 400–411. [Google Scholar] [CrossRef] [PubMed]
  529. Fitzner, D.; Schnaars, M.; van Rossum, D.; Krishnamoorthy, G.; Dibaj, P.; Bakhti, M.; Regen, T.; Hanisch, U.K.; Simons, M. Selective transfer of exosomes from oligodendrocytes to microglia by macropinocytosis. J. Cell Sci. 2011, 124, 447–458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  530. Chang, C.; Lang, H.; Geng, N.; Wang, J.; Li, N.; Wang, X. Exosomes of BV-2 cells induced by alpha-synuclein: Important mediator of neurodegeneration in PD. Neurosci. Lett. 2013, 548, 190–195. [Google Scholar] [CrossRef]
  531. Guo, M.; Wang, J.; Zhao, Y.; Feng, Y.; Han, S.; Dong, Q.; Cui, M.; Tieu, K. Microglial exosomes facilitate alpha-synuclein transmission in Parkinson’s disease. Brain 2020, 143, 1476–1497. [Google Scholar] [CrossRef]
  532. George, S.; Rey, N.L.; Tyson, T.; Esquibel, C.; Meyerdirk, L.; Schulz, E.; Pierce, S.; Burmeister, A.R.; Madaj, Z.; Steiner, J.A.; et al. Microglia affect alpha-synuclein cell-to-cell transfer in a mouse model of Parkinson’s disease. Mol. Neurodegener. 2019, 14, 34. [Google Scholar] [CrossRef] [PubMed]
  533. Deitmer, J.W.; Theparambil, S.M.; Ruminot, I.; Noor, S.I.; Becker, H.M. Energy dynamics in the brain: Contributions of astrocytes to metabolism and pH homeostasis. Front. Neurosci. 2019, 13, 1301. [Google Scholar] [CrossRef]
  534. Dienel, G.A.; Carlson, G.M. Major advances in brain glycogen research: Understanding of the roles of glycogen Have evolved from emergency fuel reserve to dynamic, regulated participant in diverse brain functions. Adv. Neurobiol. 2019, 23, 1–16. [Google Scholar] [CrossRef]
  535. Halassa, M.M.; Haydon, P.G. Integrated brain circuits: Astrocytic networks modulate neuronal activity and behavior. Annu. Rev. Physiol. 2010, 72, 335–355. [Google Scholar] [CrossRef] [Green Version]
  536. Haydon, P.G.; Carmignoto, G. Astrocyte control of synaptic transmission and neurovascular coupling. Physiol. Rev. 2006, 86, 1009–1031. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  537. Kimelberg, H.K.; Nedergaard, M. Functions of astrocytes and their potential as therapeutic targets. Neurotherapeutics 2010, 7, 338–353. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  538. Seth, P.; Koul, N. Astrocyte, the star avatar: Redefined. J. Biosci. 2008, 33, 405–421. [Google Scholar] [CrossRef]
  539. Vasile, F.; Dossi, E.; Rouach, N. Human astrocytes: Structure and functions in the healthy brain. Brain Struct. Funct. 2017, 222, 2017–2029. [Google Scholar] [CrossRef] [Green Version]
  540. Simard, M.; Nedergaard, M. The neurobiology of glia in the context of water and ion homeostasis. Neuroscience 2004, 129, 877–896. [Google Scholar] [CrossRef] [PubMed]
  541. Mulder, M. Sterols in the central nervous system. Curr. Opin. Clin. Nutr. Metab. Care 2009, 12, 152–158. [Google Scholar] [CrossRef]
  542. Kiray, H.; Lindsay, S.L.; Hosseinzadeh, S.; Barnett, S.C. The multifaceted role of astrocytes in regulating myelination. Exp. Neurol. 2016, 283, 541–549. [Google Scholar] [CrossRef] [Green Version]
  543. Traiffort, E.; Kassoussi, A.; Zahaf, A.; Laouarem, Y. Astrocytes and microglia as major players of myelin production in normal and pathological conditions. Front. Cell. Neurosci. 2020, 14, 79. [Google Scholar] [CrossRef] [Green Version]
  544. Halassa, M.M.; Fellin, T.; Haydon, P.G. The tripartite synapse: Roles for gliotransmission in health and disease. Trends Mol. Med. 2007, 13, 54–63. [Google Scholar] [CrossRef]
  545. Harada, K.; Kamiya, T.; Tsuboi, T. Gliotransmitter release from astrocytes: Functional, developmental, and pathological implications in the brain. Front. Neurosci. 2015, 9, 499. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  546. Agulhon, C.; Sun, M.Y.; Murphy, T.; Myers, T.; Lauderdale, K.; Fiacco, T.A. Calcium signaling and gliotransmission in normal vs. reactive astrocytes. Front. Pharmacol. 2012, 3, 139. [Google Scholar] [CrossRef] [Green Version]
  547. Tabata, H. Diverse subtypes of astrocytes and their development during corticogenesis. Front. Neurosci. 2015, 9, 114. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  548. Oberheim, N.A.; Takano, T.; Han, X.; He, W.; Lin, J.H.; Wang, F.; Xu, Q.; Wyatt, J.D.; Pilcher, W.; Ojemann, J.G.; et al. Uniquely hominid features of adult human astrocytes. J. Neurosci. 2009, 29, 3276–3287. [Google Scholar] [CrossRef] [PubMed]
  549. Sofroniew, M.V.; Vinters, H.V. Astrocytes: Biology and pathology. Acta Neuropathol. 2010, 119, 7–35. [Google Scholar] [CrossRef] [Green Version]
  550. Lee, H.J.; Suk, J.E.; Patrick, C.; Bae, E.J.; Cho, J.H.; Rho, S.; Hwang, D.; Masliah, E.; Lee, S.J. Direct transfer of alpha-synuclein from neuron to astroglia causes inflammatory responses in synucleinopathies. J. Biol. Chem. 2010, 285, 9262–9272. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  551. Saavedra, A.; Baltazar, G.; Santos, P.; Carvalho, C.M.; Duarte, E.P. Selective injury to dopaminergic neurons up-regulates GDNF in substantia nigra postnatal cell cultures: Role of neuron-glia crosstalk. Neurobiol. Dis. 2006, 23, 533–542. [Google Scholar] [CrossRef] [PubMed]
  552. Sandhu, J.K.; Gardaneh, M.; Iwasiow, R.; Lanthier, P.; Gangaraju, S.; Ribecco-Lutkiewicz, M.; Tremblay, R.; Kiuchi, K.; Sikorska, M. Astrocyte-secreted GDNF and glutathione antioxidant system protect neurons against 6OHDA cytotoxicity. Neurobiol. Dis. 2009, 33, 405–414. [Google Scholar] [CrossRef]
  553. Mori, F.; Tanji, K.; Yoshimoto, M.; Takahashi, H.; Wakabayashi, K. Demonstration of alpha-synuclein immunoreactivity in neuronal and glial cytoplasm in normal human brain tissue using proteinase K and formic acid pretreatment. Exp. Neurol. 2002, 176, 98–104. [Google Scholar] [CrossRef]
  554. Tanji, K.; Imaizumi, T.; Yoshida, H.; Mori, F.; Yoshimoto, M.; Satoh, K.; Wakabayashi, K. Expression of alpha-synuclein in a human glioma cell line and its up-regulation by interleukin-1beta. Neuroreport 2001, 12, 1909–1912. [Google Scholar] [CrossRef]
  555. Cheng, S.Y.; Trombetta, L.D. The induction of amyloid precursor protein and alpha-synuclein in rat hippocampal astrocytes by diethyldithiocarbamate and copper with or without glutathione. Toxicol. Lett. 2004, 146, 139–149. [Google Scholar] [CrossRef]
  556. Stevenson, T.J.; Murray, H.C.; Turner, C.; Faull, R.L.M.; Dieriks, B.V.; Curtis, M.A. alpha-synuclein inclusions are abundant in non-neuronal cells in the anterior olfactory nucleus of the Parkinson’s disease olfactory bulb. Sci. Rep. 2020, 10, 6682. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  557. Sorrentino, Z.A.; Giasson, B.I.; Chakrabarty, P. alpha-Synuclein and astrocytes: Tracing the pathways from homeostasis to neurodegeneration in Lewy body disease. Acta Neuropathol. 2019, 138, 1–21. [Google Scholar] [CrossRef] [PubMed]
  558. Piao, Y.S.; Mori, F.; Hayashi, S.; Tanji, K.; Yoshimoto, M.; Kakita, A.; Wakabayashi, K.; Takahashi, H. Alpha-synuclein pathology affecting Bergmann glia of the cerebellum in patients with alpha-synucleinopathies. Acta Neuropathol. 2003, 105, 403–409. [Google Scholar] [CrossRef]
  559. Hishikawa, N.; Hashizume, Y.; Yoshida, M.; Sobue, G. Widespread occurrence of argyrophilic glial inclusions in Parkinson’s disease. Neuropathol. Appl. Neurobiol. 2001, 27, 362–372. [Google Scholar] [CrossRef]
  560. Terada, S.; Ishizu, H.; Yokota, O.; Tsuchiya, K.; Nakashima, H.; Ishihara, T.; Fujita, D.; Ueda, K.; Ikeda, K.; Kuroda, S. Glial involvement in diffuse Lewy body disease. Acta Neuropathol. 2003, 105, 163–169. [Google Scholar] [CrossRef]
  561. Sorrentino, Z.A.; Goodwin, M.S.; Riffe, C.J.; Dhillon, J.S.; Xia, Y.; Gorion, K.M.; Vijayaraghavan, N.; McFarland, K.N.; Golbe, L.I.; Yachnis, A.T.; et al. Unique alpha-synuclein pathology within the amygdala in Lewy body dementia: Implications for disease initiation and progression. Acta Neuropathol. Commun. 2019, 7, 142. [Google Scholar] [CrossRef] [Green Version]
  562. Booth, H.D.E.; Hirst, W.D.; Wade-Martins, R. The role of astrocyte dysfunction in Parkinson’s disease pathogenesis. Trends Neurosci. 2017, 40, 358–370. [Google Scholar] [CrossRef] [Green Version]
  563. Ding, Z.B.; Song, L.J.; Wang, Q.; Kumar, G.; Yan, Y.Q.; Ma, C.G. Astrocytes: A double-edged sword in neurodegenerative diseases. Neural. Regen. Res. 2021, 16, 1702–1710. [Google Scholar] [CrossRef]
  564. Ghiglieri, V.; Calabrese, V.; Calabresi, P. Alpha-synuclein: From early synaptic dysfunction to neurodegeneration. Front. Neurol. 2018, 9, 295. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  565. Sacino, A.N.; Brooks, M.M.; Chakrabarty, P.; Saha, K.; Khoshbouei, H.; Golde, T.E.; Giasson, B.I. Proteolysis of alpha-synuclein fibrils in the lysosomal pathway limits induction of inclusion pathology. J. Neurochem. 2017, 140, 662–678. [Google Scholar] [CrossRef] [Green Version]
  566. Erustes, A.G.; Stefani, F.Y.; Terashima, J.Y.; Stilhano, R.S.; Monteforte, P.T.; da Silva Pereira, G.J.; Han, S.W.; Calgarotto, A.K.; Hsu, Y.T.; Ureshino, R.P.; et al. Overexpression of alpha-synuclein in an astrocyte cell line promotes autophagy inhibition and apoptosis. J. Neurosci. Res. 2018, 96, 160–171. [Google Scholar] [CrossRef] [PubMed]
  567. Knott, C.; Stern, G.; Kingsbury, A.; Welcher, A.A.; Wilkin, G.P. Elevated glial brain-derived neurotrophic factor in Parkinson’s diseased nigra. Parkinsonism Relat. Disord. 2002, 8, 329–341. [Google Scholar] [CrossRef]
  568. Damier, P.; Hirsch, E.C.; Zhang, P.; Agid, Y.; Javoy-Agid, F. Glutathione peroxidase, glial cells and Parkinson’s disease. Neuroscience 1993, 52, 1–6. [Google Scholar] [CrossRef]
  569. Mythri, R.B.; Venkateshappa, C.; Harish, G.; Mahadevan, A.; Muthane, U.B.; Yasha, T.C.; Srinivas Bharath, M.M.; Shankar, S.K. Evaluation of markers of oxidative stress, antioxidant function and astrocytic proliferation in the striatum and frontal cortex of Parkinson’s disease brains. Neurochem. Res. 2011, 36, 1452–1463. [Google Scholar] [CrossRef]
  570. Hua, J.; Yin, N.; Xu, S.; Chen, Q.; Tao, T.; Zhang, J.; Ding, J.; Fan, Y.; Hu, G. Enhancing the astrocytic clearance of extracellular alpha-synuclein aggregates by ginkgolides attenuates neural cell injury. Cell Mol. Neurobiol. 2019, 39, 1017–1028. [Google Scholar] [CrossRef] [PubMed]
  571. Loria, F.; Vargas, J.Y.; Bousset, L.; Syan, S.; Salles, A.; Melki, R.; Zurzolo, C. alpha-Synuclein transfer between neurons and astrocytes indicates that astrocytes play a role in degradation rather than in spreading. Acta Neuropathol. 2017, 134, 789–808. [Google Scholar] [CrossRef]
  572. Di Domenico, A.; Carola, G.; Calatayud, C.; Pons-Espinal, M.; Munoz, J.P.; Richaud-Patin, Y.; Fernandez-Carasa, I.; Gut, M.; Faella, A.; Parameswaran, J.; et al. Patient-specific iPSC-derived astrocytes contribute to non-cell-autonomous neurodegeneration in Parkinson’s disease. Stem Cell Rep. 2019, 12, 213–229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  573. Sacino, A.N.; Thomas, M.A.; Ceballos-Diaz, C.; Cruz, P.E.; Rosario, A.M.; Lewis, J.; Giasson, B.I.; Golde, T.E. Conformational templating of alpha-synuclein aggregates in neuronal-glial cultures. Mol. Neurodegener. 2013, 8, 17. [Google Scholar] [CrossRef] [Green Version]
  574. Cavaliere, F.; Cerf, L.; Dehay, B.; Ramos-Gonzalez, P.; De Giorgi, F.; Bourdenx, M.; Bessede, A.; Obeso, J.A.; Matute, C.; Ichas, F.; et al. In vitro alpha-synuclein neurotoxicity and spreading among neurons and astrocytes using Lewy body extracts from Parkinson disease brains. Neurobiol. Dis. 2017, 103, 101–112. [Google Scholar] [CrossRef] [PubMed]
  575. Sorrentino, Z.A.; Xia, Y.; Funk, C.; Riffe, C.J.; Rutherford, N.J.; Ceballos Diaz, C.; Sacino, A.N.; Price, N.D.; Golde, T.E.; Giasson, B.I.; et al. Motor neuron loss and neuroinflammation in a model of alpha-synuclein-induced neurodegeneration. Neurobiol. Dis. 2018, 120, 98–106. [Google Scholar] [CrossRef]
  576. Sorrentino, Z.A.; Brooks, M.M.T.; Hudson, V., 3rd; Rutherford, N.J.; Golde, T.E.; Giasson, B.I.; Chakrabarty, P. Intrastriatal injection of alpha-synuclein can lead to widespread synucleinopathy independent of neuroanatomic connectivity. Mol. Neurodegener. 2017, 12, 40. [Google Scholar] [CrossRef]
  577. Wang, Y.; Cui, J.; Sun, X.; Zhang, Y. Tunneling-nanotube development in astrocytes depends on p53 activation. Cell Death Differ. 2011, 18, 732–742. [Google Scholar] [CrossRef] [Green Version]
  578. Rostami, J.; Holmqvist, S.; Lindstrom, V.; Sigvardson, J.; Westermark, G.T.; Ingelsson, M.; Bergstrom, J.; Roybon, L.; Erlandsson, A. Human astrocytes transfer aggregated alpha-synuclein via tunneling nanotubes. J. Neurosci. 2017, 37, 11835–11853. [Google Scholar] [CrossRef] [Green Version]
  579. Jiang, C.; Hopfner, F.; Katsikoudi, A.; Hein, R.; Catli, C.; Evetts, S.; Huang, Y.; Wang, H.; Ryder, J.W.; Kuhlenbaeumer, G.; et al. Serum neuronal exosomes predict and differentiate Parkinson’s disease from atypical parkinsonism. J. Neurol. Neurosurg. Psychiatry 2020, 91, 720–729. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  580. Karampetsou, M.; Sykioti, V.S.; Leandrou, E.; Melachroinou, K.; Lambiris, A.; Giannelos, A.; Emmanouilidou, E.; Vekrellis, K. Intrastriatal administration of exosome-associated pathological alpha-synuclein is not sufficient by itself to cause pathology transmission. Front. Neurosci 2020, 14, 246. [Google Scholar] [CrossRef] [Green Version]
  581. Fan, R.Z.; Guo, M.; Luo, S.; Cui, M.; Tieu, K. Exosome release and neuropathology induced by alpha-synuclein: New insights into protective mechanisms of Drp1 inhibition. Acta Neuropathol. Commun. 2019, 7, 184. [Google Scholar] [CrossRef]
  582. Niu, M.; Li, Y.; Li, G.; Zhou, L.; Luo, N.; Yao, M.; Kang, W.; Liu, J. A longitudinal study on alpha-synuclein in plasma neuronal exosomes as a biomarker for Parkinson’s disease development and progression. Eur. J. Neurol. 2020, 27, 967–974. [Google Scholar] [CrossRef]
  583. Pascua-Maestro, R.; Gonzalez, E.; Lillo, C.; Ganfornina, M.D.; Falcon-Perez, J.M.; Sanchez, D. Extracellular vesicles secreted by astroglial cells transport apolipoprotein D to neurons and mediate neuronal survival upon oxidative stress. Front. Cell Neurosci. 2018, 12, 526. [Google Scholar] [CrossRef]
  584. Shakespear, N.; Ogura, M.; Yamaki, J.; Homma, Y. Astrocyte-Derived Exosomal microRNA miR-200a-3p Prevents MPP(+)-induced apoptotic cell death through down-regulation of MKK4. Neurochem. Res. 2020, 45, 1020–1033. [Google Scholar] [CrossRef] [PubMed]
  585. Meng, Y.; Ding, J.; Li, C.; Fan, H.; He, Y.; Qiu, P. Transfer of pathological alpha-synuclein from neurons to astrocytes via exosomes causes inflammatory responses after METH exposure. Toxicol. Lett. 2020, 331, 188–199. [Google Scholar] [CrossRef]
  586. Ngolab, J.; Trinh, I.; Rockenstein, E.; Mante, M.; Florio, J.; Trejo, M.; Masliah, D.; Adame, A.; Masliah, E.; Rissman, R.A. Brain-derived exosomes from dementia with Lewy bodies propagate alpha-synuclein pathology. Acta Neuropathol. Commun. 2017, 5, 46. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  587. Klegeris, A.; Giasson, B.I.; Zhang, H.; Maguire, J.; Pelech, S.; McGeer, P.L. Alpha-synuclein and its disease-causing mutants induce ICAM-1 and IL-6 in human astrocytes and astrocytoma cells. FASEB J. 2006, 20, 2000–2008. [Google Scholar] [CrossRef] [PubMed]
  588. Rannikko, E.H.; Weber, S.S.; Kahle, P.J. Exogenous alpha-synuclein induces toll-like receptor 4 dependent inflammatory responses in astrocytes. BMC Neurosci. 2015, 16, 57. [Google Scholar] [CrossRef] [Green Version]
  589. Angelova, P.R.; Ludtmann, M.H.; Horrocks, M.H.; Negoda, A.; Cremades, N.; Klenerman, D.; Dobson, C.M.; Wood, N.W.; Pavlov, E.V.; Gandhi, S.; et al. Ca2+ is a key factor in alpha-synuclein-induced neurotoxicity. J. Cell Sci. 2016, 129, 1792–1801. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  590. Cremades, N.; Cohen, S.I.; Deas, E.; Abramov, A.Y.; Chen, A.Y.; Orte, A.; Sandal, M.; Clarke, R.W.; Dunne, P.; Aprile, F.A.; et al. Direct observation of the interconversion of normal and toxic forms of alpha-synuclein. Cell 2012, 149, 1048–1059. [Google Scholar] [CrossRef] [Green Version]
  591. Stefanova, N.; Klimaschewski, L.; Poewe, W.; Wenning, G.K.; Reindl, M. Glial cell death induced by overexpression of alpha-synuclein. J. Neurosci. Res. 2001, 65, 432–438. [Google Scholar] [CrossRef]
  592. Yun, S.P.; Kam, T.I.; Panicker, N.; Kim, S.; Oh, Y.; Park, J.S.; Kwon, S.H.; Park, Y.J.; Karuppagounder, S.S.; Park, H.; et al. Block of A1 astrocyte conversion by microglia is neuroprotective in models of Parkinson’s disease. Nat. Med. 2018, 24, 931–938. [Google Scholar] [CrossRef]
  593. Gu, X.L.; Long, C.X.; Sun, L.; Xie, C.; Lin, X.; Cai, H. Astrocytic expression of Parkinson’s disease-related A53T alpha-synuclein causes neurodegeneration in mice. Mol. Brain 2010, 3, 12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  594. Koob, A.O.; Paulino, A.D.; Masliah, E. GFAP reactivity, apolipoprotein E redistribution and cholesterol reduction in human astrocytes treated with alpha-synuclein. Neurosci Lett 2010, 469, 11–14. [Google Scholar] [CrossRef] [Green Version]
  595. Chavarria, C.; Rodriguez-Bottero, S.; Quijano, C.; Cassina, P.; Souza, J.M. Impact of monomeric, oligomeric and fibrillar alpha-synuclein on astrocyte reactivity and toxicity to neurons. Biochem. J. 2018, 475, 3153–3169. [Google Scholar] [CrossRef]
  596. Choi, D.K.; Pennathur, S.; Perier, C.; Tieu, K.; Teismann, P.; Wu, D.C.; Jackson-Lewis, V.; Vila, M.; Vonsattel, J.P.; Heinecke, J.W.; et al. Ablation of the inflammatory enzyme myeloperoxidase mitigates features of Parkinson’s disease in mice. J. Neurosci. 2005, 25, 6594–6600. [Google Scholar] [CrossRef] [PubMed]
  597. Hashioka, S.; Klegeris, A.; Schwab, C.; Yu, S.; McGeer, P.L. Differential expression of interferon-gamma receptor on human glial cells in vivo and in vitro. J. Neuroimmunol. 2010, 225, 91–99. [Google Scholar] [CrossRef]
  598. Barcia, C.; Ros, C.M.; Annese, V.; Gomez, A.; Ros-Bernal, F.; Aguado-Yera, D.; Martinez-Pagan, M.E.; de Pablos, V.; Fernandez-Villalba, E.; Herrero, M.T. IFN-gamma signaling, with the synergistic contribution of TNF-alpha, mediates cell specific microglial and astroglial activation in experimental models of Parkinson’s disease. Cell Death Dis. 2011, 2, e142. [Google Scholar] [CrossRef] [Green Version]
  599. Diaz, E.F.; Labra, V.C.; Alvear, T.F.; Mellado, L.A.; Inostroza, C.A.; Oyarzun, J.E.; Salgado, N.; Quintanilla, R.A.; Orellana, J.A. Connexin 43 hemichannels and pannexin-1 channels contribute to the alpha-synuclein-induced dysfunction and death of astrocytes. Glia 2019, 67, 1598–1619. [Google Scholar] [CrossRef]
  600. Sheng, L.; Stewart, T.; Yang, D.; Thorland, E.; Soltys, D.; Aro, P.; Khrisat, T.; Xie, Z.; Li, N.; Liu, Z.; et al. Erythrocytic alpha-synuclein contained in microvesicles regulates astrocytic glutamate homeostasis: A new perspective on Parkinson’s disease pathogenesis. Acta Neuropathol. Commun. 2020, 8, 102. [Google Scholar] [CrossRef] [PubMed]
  601. Liu, M.; Qin, L.; Wang, L.; Tan, J.; Zhang, H.; Tang, J.; Shen, X.; Tan, L.; Wang, C. alphasynuclein induces apoptosis of astrocytes by causing dysfunction of the endoplasmic reticulumGolgi compartment. Mol. Med. Rep. 2018, 18, 322–332. [Google Scholar] [CrossRef] [Green Version]
  602. Lindstrom, V.; Gustafsson, G.; Sanders, L.H.; Howlett, E.H.; Sigvardson, J.; Kasrayan, A.; Ingelsson, M.; Bergstrom, J.; Erlandsson, A. Extensive uptake of alpha-synuclein oligomers in astrocytes results in sustained intracellular deposits and mitochondrial damage. Mol. Cell Neurosci. 2017, 82, 143–156. [Google Scholar] [CrossRef]
  603. Jo, M.; Kim, J.H.; Song, G.J.; Seo, M.; Hwang, E.M.; Suk, K. Astrocytic Orosomucoid-2 Modulates Microglial Activation and Neuroinflammation. J. Neurosci. 2017, 37, 2878–2894. [Google Scholar] [CrossRef] [Green Version]
  604. Rohl, C.; Sievers, J. Microglia is activated by astrocytes in trimethyltin intoxication. Toxicol. Appl. Pharmacol. 2005, 204, 36–45. [Google Scholar] [CrossRef] [PubMed]
  605. Lee, H.J.; Kim, C.; Lee, S.J. Alpha-synuclein stimulation of astrocytes: Potential role for neuroinflammation and neuroprotection. Oxid. Med. Cell Longev. 2010, 3, 283–287. [Google Scholar] [CrossRef] [PubMed]
  606. Farina, C.; Aloisi, F.; Meinl, E. Astrocytes are active players in cerebral innate immunity. Trends Immunol. 2007, 28, 138–145. [Google Scholar] [CrossRef] [PubMed]
  607. Di Marco Vieira, B.; Radford, R.A.W.; Hayashi, J.; Eaton, E.D.; Greenaway, B.; Jambas, M.; Petcu, E.B.; Chung, R.S.; Pountney, D.L. Extracellular alpha-synuclein promotes a neuroinhibitory secretory phenotype in astrocytes. Life 2020, 10, 183. [Google Scholar] [CrossRef]
  608. Miklossy, J.; Doudet, D.D.; Schwab, C.; Yu, S.; McGeer, E.G.; McGeer, P.L. Role of ICAM-1 in persisting inflammation in Parkinson disease and MPTP monkeys. Exp. Neurol. 2006, 197, 275–283. [Google Scholar] [CrossRef]
  609. Barry, M.; Bradford, L.I.M.; David, A.; Hume, C.P.; Neil, A. Mabbott. Complete microglia deficiency accelerates prion disease without enhancing CNS prion accumulation. Cell Rep. 2021. [Google Scholar] [CrossRef]
  610. Watanabe, H.; Saito, Y.; Terao, S.; Ando, T.; Kachi, T.; Mukai, E.; Aiba, I.; Abe, Y.; Tamakoshi, A.; Doyu, M.; et al. Progression and prognosis in multiple system atrophy: An analysis of 230 Japanese patients. Brain 2002, 125, 1070–1083. [Google Scholar] [CrossRef]
  611. Ozawa, T.; Paviour, D.; Quinn, N.P.; Josephs, K.A.; Sangha, H.; Kilford, L.; Healy, D.G.; Wood, N.W.; Lees, A.J.; Holton, J.L.; et al. The spectrum of pathological involvement of the striatonigral and olivopontocerebellar systems in multiple system atrophy: Clinicopathological correlations. Brain 2004, 127, 2657–2671. [Google Scholar] [CrossRef] [Green Version]
  612. Jellinger, K.A.; Seppi, K.; Wenning, G.K. Grading of neuropathology in multiple system atrophy: Proposal for a novel scale. Mov. Disord. 2005, 20 (Suppl. S12), S29–S36. [Google Scholar] [CrossRef]
  613. Radford, R.; Rcom-H’cheo-Gauthier, A.; Wong, M.B.; Eaton, E.D.; Quilty, M.; Blizzard, C.; Norazit, A.; Meedeniya, A.; Vickers, J.C.; Gai, W.P.; et al. The degree of astrocyte activation in multiple system atrophy is inversely proportional to the distance to alpha-synuclein inclusions. Mol. Cell Neurosci. 2015, 65, 68–81. [Google Scholar] [CrossRef] [PubMed]
  614. Nakamura, K.; Mori, F.; Kon, T.; Tanji, K.; Miki, Y.; Tomiyama, M.; Kurotaki, H.; Toyoshima, Y.; Kakita, A.; Takahashi, H.; et al. Accumulation of phosphorylated alpha-synuclein in subpial and periventricular astrocytes in multiple system atrophy of long duration. Neuropathology 2016, 36, 157–167. [Google Scholar] [CrossRef] [PubMed]
  615. Shults, C.W.; Rockenstein, E.; Crews, L.; Adame, A.; Mante, M.; Larrea, G.; Hashimoto, M.; Song, D.; Iwatsubo, T.; Tsuboi, K.; et al. Neurological and neurodegenerative alterations in a transgenic mouse model expressing human alpha-synuclein under oligodendrocyte promoter: Implications for multiple system atrophy. J. Neurosci. 2005, 25, 10689–10699. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  616. Valera, E.; Ubhi, K.; Mante, M.; Rockenstein, E.; Masliah, E. Antidepressants reduce neuroinflammatory responses and astroglial alpha-synuclein accumulation in a transgenic mouse model of multiple system atrophy. Glia 2014, 62, 317–337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  617. Yazawa, I.; Giasson, B.I.; Sasaki, R.; Zhang, B.; Joyce, S.; Uryu, K.; Trojanowski, J.Q.; Lee, V.M. Mouse model of multiple system atrophy alpha-synuclein expression in oligodendrocytes causes glial and neuronal degeneration. Neuron 2005, 45, 847–859. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  618. Stefanova, N.; Reindl, M.; Neumann, M.; Haass, C.; Poewe, W.; Kahle, P.J.; Wenning, G.K. Oxidative stress in transgenic mice with oligodendroglial alpha-synuclein overexpression replicates the characteristic neuropathology of multiple system atrophy. Am. J. Pathol. 2005, 166, 869–876. [Google Scholar] [CrossRef]
  619. Bradl, M.; Lassmann, H. Oligodendrocytes: Biology and pathology. Acta Neuropathol. 2010, 119, 37–53. [Google Scholar] [CrossRef] [Green Version]
  620. Wilkins, A.; Majed, H.; Layfield, R.; Compston, A.; Chandran, S. Oligodendrocytes promote neuronal survival and axonal length by distinct intracellular mechanisms: A novel role for oligodendrocyte-derived glial cell line-derived neurotrophic factor. J. Neurosci. 2003, 23, 4967–4974. [Google Scholar] [CrossRef] [Green Version]
  621. Bean, B.P. The action potential in mammalian central neurons. Nat. Rev. Neurosci. 2007, 8, 451–465. [Google Scholar] [CrossRef]
  622. Kuhn, S.; Gritti, L.; Crooks, D.; Dombrowski, Y. Oligodendrocytes in development, myelin generation and beyond. Cells 2019, 8, 1424. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  623. Simons, M.; Nave, K.A. Oligodendrocytes: Myelination and axonal support. Cold Spring Harb. Perspect. Biol. 2015, 8, a020479. [Google Scholar] [CrossRef] [PubMed]
  624. Hofmann, K.; Rodriguez-Rodriguez, R.; Gaebler, A.; Casals, N.; Scheller, A.; Kuerschner, L. Astrocytes and oligodendrocytes in grey and white matter regions of the brain metabolize fatty acids. Sci. Rep. 2017, 7, 10779. [Google Scholar] [CrossRef]
  625. Dulamea, A.O. Role of oligodendrocyte dysfunction in demyelination, remyelination and neurodegeneration in multiple sclerosis. Adv. Exp. Med. Biol. 2017, 958, 91–127. [Google Scholar] [CrossRef]
  626. Wong, J.H.; Halliday, G.M.; Kim, W.S. Exploring myelin dysfunction in multiple system atrophy. Exp. Neurobiol. 2014, 23, 337–344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  627. Fellner, L.; Jellinger, K.A.; Wenning, G.K.; Stefanova, N. Glial dysfunction in the pathogenesis of alpha-synucleinopathies: Emerging concepts. Acta Neuropathol. 2011, 121, 675–693. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  628. Campbell, B.C.; McLean, C.A.; Culvenor, J.G.; Gai, W.P.; Blumbergs, P.C.; Jakala, P.; Beyreuther, K.; Masters, C.L.; Li, Q.X. The solubility of alpha-synuclein in multiple system atrophy differs from that of dementia with Lewy bodies and Parkinson’s disease. J. Neurochem. 2001, 76, 87–96. [Google Scholar] [CrossRef]
  629. Yamada, T.; McGeer, P.L.; McGeer, E.G. Lewy bodies in Parkinson’s disease are recognized by antibodies to complement proteins. Acta Neuropathol. 1992, 84, 100–104. [Google Scholar] [CrossRef]
  630. Yamada, T.; McGeer, P.L.; McGeer, E.G. Some immunohistochemical features of argyrophilic grain dementia with normal cortical choline acetyltransferase levels but extensive subcortical pathology and markedly reduced dopamine. J. Geriatr. Psychiatry Neurol. 1992, 5, 3–13. [Google Scholar] [CrossRef] [PubMed]
  631. Braak, H.; Del Tredici, K. Poor and protracted myelination as a contributory factor to neurodegenerative disorders. Neurobiol. Aging 2004, 25, 19–23. [Google Scholar] [CrossRef] [PubMed]
  632. Braak, H.; Del Tredici, K. Neuroanatomy and pathology of sporadic Parkinson’s disease. Adv. Anat. Embryol. Cell Biol. 2009, 201, 1–119. [Google Scholar]
  633. Takagi, S.; Hayakawa, N.; Kimoto, H.; Kato, H.; Araki, T. Damage to oligodendrocytes in the striatum after MPTP neurotoxicity in mice. J. Neural. Transm. 2007, 114, 1553–1557. [Google Scholar] [CrossRef]
  634. Wenning, G.K.; Stefanova, N.; Jellinger, K.A.; Poewe, W.; Schlossmacher, M.G. Multiple system atrophy: A primary oligodendrogliopathy. Ann. Neurol. 2008, 64, 239–246. [Google Scholar] [CrossRef] [PubMed]
  635. Ozawa, T.; Okuizumi, K.; Ikeuchi, T.; Wakabayashi, K.; Takahashi, H.; Tsuji, S. Analysis of the expression level of alpha-synuclein mRNA using postmortem brain samples from pathologically confirmed cases of multiple system atrophy. Acta Neuropathol. 2001, 102, 188–190. [Google Scholar] [CrossRef] [PubMed]
  636. Papp, M.I.; Lantos, P.L. The distribution of oligodendroglial inclusions in multiple system atrophy and its relevance to clinical symptomatology. Brain 1994, 117, 235–243. [Google Scholar] [CrossRef] [PubMed]
  637. Jellinger, K.A.; Lantos, P.L. Papp-Lantos inclusions and the pathogenesis of multiple system atrophy: An update. Acta Neuropathol. 2010, 119, 657–667. [Google Scholar] [CrossRef] [PubMed]
  638. Papp, M.I.; Lantos, P.L. Accumulation of tubular structures in oligodendroglial and neuronal cells as the basic alteration in multiple system atrophy. J. Neurol. Sci. 1992, 107, 172–182. [Google Scholar] [CrossRef]
  639. Yoshida, M. Multiple system atrophy: Alpha-synuclein and neuronal degeneration. Neuropathology 2007, 27, 484–493. [Google Scholar] [CrossRef]
  640. Takeda, A.; Arai, N.; Komori, T.; Kato, S.; Oda, M. Neuronal inclusions in the dentate fascia in patients with multiple system atrophy. Neurosci. Lett. 1997, 227, 157–160. [Google Scholar] [CrossRef]
  641. Papp, M.I.; Kahn, J.E.; Lantos, P.L. Glial cytoplasmic inclusions in the CNS of patients with multiple system atrophy (striatonigral degeneration, olivopontocerebellar atrophy and Shy-Drager syndrome). J. Neurol. Sci. 1989, 94, 79–100. [Google Scholar] [CrossRef]
  642. Wakabayashi, K.; Yoshimoto, M.; Tsuji, S.; Takahashi, H. Alpha-synuclein immunoreactivity in glial cytoplasmic inclusions in multiple system atrophy. Neurosci. Lett. 1998, 249, 180–182. [Google Scholar] [CrossRef]
  643. Gai, W.P.; Pountney, D.L.; Power, J.H.; Li, Q.X.; Culvenor, J.G.; McLean, C.A.; Jensen, P.H.; Blumbergs, P.C. alpha-Synuclein fibrils constitute the central core of oligodendroglial inclusion filaments in multiple system atrophy. Exp. Neurol. 2003, 181, 68–78. [Google Scholar] [CrossRef]
  644. McCormack, A.; Chegeni, N.; Chegini, F.; Colella, A.; Power, J.; Keating, D.; Chataway, T. Purification of alpha-synuclein containing inclusions from human post mortem brain tissue. J. Neurosci. Methods 2016, 266, 141–150. [Google Scholar] [CrossRef] [PubMed]
  645. Riedel, M.; Goldbaum, O.; Richter-Landsberg, C. alpha-Synuclein promotes the recruitment of tau to protein inclusions in oligodendroglial cells: Effects of oxidative and proteolytic stress. J. Mol. Neurosci. 2009, 39, 226–234. [Google Scholar] [CrossRef]
  646. Skjoerringe, T.; Lundvig, D.M.; Jensen, P.H.; Moos, T. P25alpha/Tubulin polymerization promoting protein expression by myelinating oligodendrocytes of the developing rat brain. J. Neurochem. 2006, 99, 333–342. [Google Scholar] [CrossRef]
  647. Takahashi, M.; Tomizawa, K.; Fujita, S.C.; Sato, K.; Uchida, T.; Imahori, K. A brain-specific protein p25 is localized and associated with oligodendrocytes, neuropil, and fiber-like structures of the CA3 hippocampal region in the rat brain. J. Neurochem. 1993, 60, 228–235. [Google Scholar] [CrossRef]
  648. Otzen, D.E.; Lundvig, D.M.; Wimmer, R.; Nielsen, L.H.; Pedersen, J.R.; Jensen, P.H. p25alpha is flexible but natively folded and binds tubulin with oligomeric stoichiometry. Protein Sci. 2005, 14, 1396–1409. [Google Scholar] [CrossRef]
  649. Lindersson, E.; Lundvig, D.; Petersen, C.; Madsen, P.; Nyengaard, J.R.; Hojrup, P.; Moos, T.; Otzen, D.; Gai, W.P.; Blumbergs, P.C.; et al. p25alpha Stimulates alpha-synuclein aggregation and is co-localized with aggregated alpha-synuclein in alpha-synucleinopathies. J. Biol. Chem. 2005, 280, 5703–5715. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  650. Ejlerskov, P.; Rasmussen, I.; Nielsen, T.T.; Bergstrom, A.L.; Tohyama, Y.; Jensen, P.H.; Vilhardt, F. Tubulin polymerization-promoting protein (TPPP/p25alpha) promotes unconventional secretion of alpha-synuclein through exophagy by impairing autophagosome-lysosome fusion. J. Biol. Chem. 2013, 288, 17313–17335. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  651. Kragh, C.L.; Fillon, G.; Gysbers, A.; Hansen, H.D.; Neumann, M.; Richter-Landsberg, C.; Haass, C.; Zalc, B.; Lubetzki, C.; Gai, W.P.; et al. FAS-dependent cell death in alpha-synuclein transgenic oligodendrocyte models of multiple system atrophy. PLoS ONE 2013, 8, e55243. [Google Scholar] [CrossRef]
  652. Kragh, C.L.; Lund, L.B.; Febbraro, F.; Hansen, H.D.; Gai, W.P.; El-Agnaf, O.; Richter-Landsberg, C.; Jensen, P.H. Alpha-synuclein aggregation and Ser-129 phosphorylation-dependent cell death in oligodendroglial cells. J. Biol. Chem. 2009, 284, 10211–10222. [Google Scholar] [CrossRef] [Green Version]
  653. Mavroeidi, P.; Arvanitaki, F.; Karakitsou, A.K.; Vetsi, M.; Kloukina, I.; Zweckstetter, M.; Giller, K.; Becker, S.; Sorrentino, Z.A.; Giasson, B.I.; et al. Endogenous oligodendroglial alpha-synuclein and TPPP/p25alpha orchestrate alpha-synuclein pathology in experimental multiple system atrophy models. Acta Neuropathol. 2019, 138, 415–441. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  654. Ubhi, K.; Rockenstein, E.; Mante, M.; Inglis, C.; Adame, A.; Patrick, C.; Whitney, K.; Masliah, E. Neurodegeneration in a transgenic mouse model of multiple system atrophy is associated with altered expression of oligodendroglial-derived neurotrophic factors. J. Neurosci. 2010, 30, 6236–6246. [Google Scholar] [CrossRef] [Green Version]
  655. Wakabayashi, K.; Takahashi, H. Cellular pathology in multiple system atrophy. Neuropathology 2006, 26, 338–345. [Google Scholar] [CrossRef]
  656. Djelloul, M.; Holmqvist, S.; Boza-Serrano, A.; Azevedo, C.; Yeung, M.S.; Goldwurm, S.; Frisen, J.; Deierborg, T.; Roybon, L. Alpha-synuclein expression in the oligodendrocyte lineage: An in vitro and in vivo study using rodent and human models. Stem Cell Rep. 2015, 5, 174–184. [Google Scholar] [CrossRef] [Green Version]
  657. Asi, Y.T.; Simpson, J.E.; Heath, P.R.; Wharton, S.B.; Lees, A.J.; Revesz, T.; Houlden, H.; Holton, J.L. Alpha-synuclein mRNA expression in oligodendrocytes in MSA. Glia 2014, 62, 964–970. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  658. Reyes, J.F.; Rey, N.L.; Bousset, L.; Melki, R.; Brundin, P.; Angot, E. Alpha-synuclein transfers from neurons to oligodendrocytes. Glia 2014, 62, 387–398. [Google Scholar] [CrossRef] [PubMed]
  659. Schwarz, L.; Goldbaum, O.; Bergmann, M.; Probst-Cousin, S.; Richter-Landsberg, C. Involvement of macroautophagy in multiple system atrophy and protein aggregate formation in oligodendrocytes. J. Mol. Neurosci. 2012, 47, 256–266. [Google Scholar] [CrossRef]
  660. Rockenstein, E.; Ubhi, K.; Inglis, C.; Mante, M.; Patrick, C.; Adame, A.; Masliah, E. Neuronal to oligodendroglial alpha-synuclein redistribution in a double transgenic model of multiple system atrophy. Neuroreport 2012, 23, 259–264. [Google Scholar] [CrossRef] [Green Version]
  661. Yoon, Y.S.; Ahn, W.J.; Ricarte, D.; Ortiz, D.; Shin, C.Y.; Lee, S.J.; Lee, H.J. Alpha-Synuclein Inclusion Formation in Human Oligodendrocytes. Biomol. Ther. 2021, 29, 83–89. [Google Scholar] [CrossRef] [PubMed]
  662. Hansen, C.; Angot, E.; Bergstrom, A.L.; Steiner, J.A.; Pieri, L.; Paul, G.; Outeiro, T.F.; Melki, R.; Kallunki, P.; Fog, K.; et al. Alpha-synuclein propagates from mouse brain to grafted dopaminergic neurons and seeds aggregation in cultured human cells. J. Clin. Investig. 2011, 121, 715–725. [Google Scholar] [CrossRef] [Green Version]
  663. Yamada, K.; Iwatsubo, T. Extracellular alpha-synuclein levels are regulated by neuronal activity. Mol. Neurodegener. 2018, 13, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  664. Kisos, H.; Pukass, K.; Ben-Hur, T.; Richter-Landsberg, C.; Sharon, R. Increased neuronal alpha-synuclein pathology associates with its accumulation in oligodendrocytes in mice modeling alpha-synucleinopathies. PLoS ONE 2012, 7, e46817. [Google Scholar] [CrossRef] [Green Version]
  665. Konno, M.; Hasegawa, T.; Baba, T.; Miura, E.; Sugeno, N.; Kikuchi, A.; Fiesel, F.C.; Sasaki, T.; Aoki, M.; Itoyama, Y.; et al. Suppression of dynamin GTPase decreases alpha-synuclein uptake by neuronal and oligodendroglial cells: A potent therapeutic target for synucleinopathy. Mol. Neurodegener. 2012, 7, 38. [Google Scholar] [CrossRef] [Green Version]
  666. Pukass, K.; Richter-Landsberg, C. Oxidative stress promotes uptake, accumulation, and oligomerization of extracellular alpha-synuclein in oligodendrocytes. J. Mol. Neurosci. 2014, 52, 339–352. [Google Scholar] [CrossRef] [PubMed]
  667. Fruhbeis, C.; Frohlich, D.; Kuo, W.P.; Amphornrat, J.; Thilemann, S.; Saab, A.S.; Kirchhoff, F.; Mobius, W.; Goebbels, S.; Nave, K.A.; et al. Neurotransmitter-triggered transfer of exosomes mediates oligodendrocyte-neuron communication. PLoS Biol. 2013, 11, e1001604. [Google Scholar] [CrossRef] [Green Version]
  668. Ghidoni, R.; Benussi, L.; Binetti, G. Exosomes: The Trojan horses of neurodegeneration. Med. Hypotheses 2008, 70, 1226–1227. [Google Scholar] [CrossRef]
  669. Nakamura, S.; Kawamoto, Y.; Nakano, S.; Akiguchi, I. Expression of the endocytosis regulatory proteins Rab5 and Rabaptin-5 in glial cytoplasmic inclusions from brains with multiple system atrophy. Clin. Neuropathol. 2000, 19, 51–56. [Google Scholar]
  670. Prusiner, S.B.; Woerman, A.L.; Mordes, D.A.; Watts, J.C.; Rampersaud, R.; Berry, D.B.; Patel, S.; Oehler, A.; Lowe, J.K.; Kravitz, S.N.; et al. Evidence for alpha-synuclein prions causing multiple system atrophy in humans with parkinsonism. Proc. Natl. Acad. Sci. USA 2015, 112, E5308–E5317. [Google Scholar] [CrossRef] [Green Version]
  671. Watts, J.C.; Giles, K.; Oehler, A.; Middleton, L.; Dexter, D.T.; Gentleman, S.M.; DeArmond, S.J.; Prusiner, S.B. Transmission of multiple system atrophy prions to transgenic mice. Proc. Natl. Acad. Sci. USA 2013, 110, 19555–19560. [Google Scholar] [CrossRef] [Green Version]
  672. Woerman, A.L.; Oehler, A.; Kazmi, S.A.; Lee, J.; Halliday, G.M.; Middleton, L.T.; Gentleman, S.M.; Mordes, D.A.; Spina, S.; Grinberg, L.T.; et al. Multiple system atrophy prions retain strain specificity after serial propagation in two different Tg(SNCA*A53T) mouse lines. Acta Neuropathol. 2019, 137, 437–454. [Google Scholar] [CrossRef]
  673. Bernis, M.E.; Babila, J.T.; Breid, S.; Wusten, K.A.; Wullner, U.; Tamguney, G. Prion-like propagation of human brain-derived alpha-synuclein in transgenic mice expressing human wild-type alpha-synuclein. Acta Neuropathol. Commun. 2015, 3, 75. [Google Scholar] [CrossRef] [Green Version]
  674. Woerman, A.L.; Stohr, J.; Aoyagi, A.; Rampersaud, R.; Krejciova, Z.; Watts, J.C.; Ohyama, T.; Patel, S.; Widjaja, K.; Oehler, A.; et al. Propagation of prions causing synucleinopathies in cultured cells. Proc. Natl. Acad. Sci. USA 2015, 112, E4949–E4958. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  675. Kaji, S.; Maki, T.; Ishimoto, T.; Yamakado, H.; Takahashi, R. Insights into the pathogenesis of multiple system atrophy: Focus on glial cytoplasmic inclusions. Transl. Neurodegener. 2020, 9, 7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  676. Reyes, J.F.; Sackmann, C.; Hoffmann, A.; Svenningsson, P.; Winkler, J.; Ingelsson, M.; Hallbeck, M. Binding of alpha-synuclein oligomers to Cx32 facilitates protein uptake and transfer in neurons and oligodendrocytes. Acta Neuropathol. 2019, 138, 23–47. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  677. Kaji, S.; Maki, T.; Kinoshita, H.; Uemura, N.; Ayaki, T.; Kawamoto, Y.; Furuta, T.; Urushitani, M.; Hasegawa, M.; Kinoshita, Y.; et al. Pathological endogenous alpha-synuclein accumulation in oligodendrocyte precursor cells potentially induces inclusions in multiple system atrophy. Stem Cell Rep. 2018, 10, 356–365. [Google Scholar] [CrossRef] [Green Version]
  678. Uemura, N.; Uemura, M.T.; Lo, A.; Bassil, F.; Zhang, B.; Luk, K.C.; Lee, V.M.; Takahashi, R.; Trojanowski, J.Q. Slow progressive accumulation of oligodendroglial alpha-synuclein (alpha-Syn) pathology in synthetic alpha-Syn Fibril-Induced Mouse Models of Synucleinopathy. J. Neuropathol Exp. Neurol 2019, 78, 877–890. [Google Scholar] [CrossRef]
  679. Ettle, B.; Reiprich, S.; Deusser, J.; Schlachetzki, J.C.; Xiang, W.; Prots, I.; Masliah, E.; Winner, B.; Wegner, M.; Winkler, J. Intracellular alpha-synuclein affects early maturation of primary oligodendrocyte progenitor cells. Mol. Cell Neurosci. 2014, 62, 68–78. [Google Scholar] [CrossRef] [Green Version]
  680. May, V.E.; Ettle, B.; Poehler, A.M.; Nuber, S.; Ubhi, K.; Rockenstein, E.; Winner, B.; Wegner, M.; Masliah, E.; Winkler, J. alpha-Synuclein impairs oligodendrocyte progenitor maturation in multiple system atrophy. Neurobiol. Aging 2014, 35, 2357–2368. [Google Scholar] [CrossRef] [Green Version]
  681. Ettle, B.; Kerman, B.E.; Valera, E.; Gillmann, C.; Schlachetzki, J.C.; Reiprich, S.; Buttner, C.; Ekici, A.B.; Reis, A.; Wegner, M.; et al. alpha-Synuclein-induced myelination deficit defines a novel interventional target for multiple system atrophy. Acta Neuropathol. 2016, 132, 59–75. [Google Scholar] [CrossRef] [Green Version]
  682. Kaji, S.; Maki, T.; Ueda, J.; Ishimoto, T.; Inoue, Y.; Yasuda, K.; Sawamura, M.; Hikawa, R.; Ayaki, T.; Yamakado, H.; et al. BCAS1-positive immature oligodendrocytes are affected by the alpha-synuclein-induced pathology of multiple system atrophy. Acta Neuropathol. Commun. 2020, 8, 120. [Google Scholar] [CrossRef]
  683. Terni, B.; Rey, M.J.; Boluda, S.; Torrejon-Escribano, B.; Sabate, M.P.; Calopa, M.; van Leeuwen, F.W.; Ferrer, I. Mutant ubiquitin and p62 immunoreactivity in cases of combined multiple system atrophy and Alzheimer’s disease. Acta Neuropathol. 2007, 113, 403–416. [Google Scholar] [CrossRef] [PubMed]
  684. Tanji, K.; Odagiri, S.; Maruyama, A.; Mori, F.; Kakita, A.; Takahashi, H.; Wakabayashi, K. Alteration of autophagosomal proteins in the brain of multiple system atrophy. Neurobiol. Dis. 2013, 49, 190–198. [Google Scholar] [CrossRef] [PubMed]
  685. Odagiri, S.; Tanji, K.; Mori, F.; Kakita, A.; Takahashi, H.; Wakabayashi, K. Autophagic adapter protein NBR1 is localized in Lewy bodies and glial cytoplasmic inclusions and is involved in aggregate formation in alpha-synucleinopathy. Acta Neuropathol. 2012, 124, 173–186. [Google Scholar] [CrossRef]
  686. Mori, F.; Nishie, M.; Piao, Y.S.; Kito, K.; Kamitani, T.; Takahashi, H.; Wakabayashi, K. Accumulation of NEDD8 in neuronal and glial inclusions of neurodegenerative disorders. Neuropathol. Appl. Neurobiol. 2005, 31, 53–61. [Google Scholar] [CrossRef] [PubMed]
  687. Miki, Y.; Mori, F.; Tanji, K.; Kakita, A.; Takahashi, H.; Wakabayashi, K. Accumulation of histone deacetylase 6, an aggresome-related protein, is specific to Lewy bodies and glial cytoplasmic inclusions. Neuropathology 2011, 31, 561–568. [Google Scholar] [CrossRef]
  688. Miki, Y.; Tanji, K.; Mori, F.; Tatara, Y.; Utsumi, J.; Sasaki, H.; Kakita, A.; Takahashi, H.; Fimia, G.M.; Wakabayashi, K. AMBRA1, a novel alpha-synuclein-binding protein, is implicated in the pathogenesis of multiple system atrophy. Brain Pathol. 2018, 28, 28–42. [Google Scholar] [CrossRef] [PubMed]
  689. Pukass, K.; Richter-Landsberg, C. Inhibition of UCH-L1 in oligodendroglial cells results in microtubule stabilization and prevents alpha-synuclein aggregate formation by activating the autophagic pathway: Implications for multiple system atrophy. Front. Cell Neurosci. 2015, 9, 163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  690. Pukass, K.; Goldbaum, O.; Richter-Landsberg, C. Mitochondrial impairment and oxidative stress compromise autophagosomal degradation of alpha-synuclein in oligodendroglial cells. J. Neurochem. 2015, 135, 194–205. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  691. Spencer, B.; Valera, E.; Rockenstein, E.; Trejo-Morales, M.; Adame, A.; Masliah, E. A brain-targeted, modified neurosin (kallikrein-6) reduces alpha-synuclein accumulation in a mouse model of multiple system atrophy. Mol. Neurodegener. 2015, 10, 48. [Google Scholar] [CrossRef] [Green Version]
  692. Kiely, A.P.; Miners, J.S.; Courtney, R.; Strand, C.; Love, S.; Holton, J.L. Exploring the putative role of kallikrein-6, calpain-1 and cathepsin-D in the proteolytic degradation of alpha-synuclein in multiple system atrophy. Neuropathol. Appl. Neurobiol. 2019, 45, 347–360. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  693. Iwata, A.; Maruyama, M.; Akagi, T.; Hashikawa, T.; Kanazawa, I.; Tsuji, S.; Nukina, N. Alpha-synuclein degradation by serine protease neurosin: Implication for pathogenesis of synucleinopathies. Hum. Mol. Genet. 2003, 12, 2625–2635. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  694. Stefanova, N.; Kaufmann, W.A.; Humpel, C.; Poewe, W.; Wenning, G.K. Systemic proteasome inhibition triggers neurodegeneration in a transgenic mouse model expressing human alpha-synuclein under oligodendrocyte promoter: Implications for multiple system atrophy. Acta Neuropathol. 2012, 124, 51–65. [Google Scholar] [CrossRef] [Green Version]
  695. Cohlberg, J.A.; Li, J.; Uversky, V.N.; Fink, A.L. Heparin and other glycosaminoglycans stimulate the formation of amyloid fibrils from alpha-synuclein in vitro. Biochemistry 2002, 41, 1502–1511. [Google Scholar] [CrossRef] [PubMed]
  696. Maiza, A.; Chantepie, S.; Vera, C.; Fifre, A.; Huynh, M.B.; Stettler, O.; Ouidja, M.O.; Papy-Garcia, D. The role of heparan sulfates in protein aggregation and their potential impact on neurodegeneration. FEBS Lett. 2018, 592, 3806–3818. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  697. Iozzo, R.V. Matrix proteoglycans: From molecular design to cellular function. Annu. Rev. Biochem. 1998, 67, 609–652. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  698. Medeiros, G.F.; Mendes, A.; Castro, R.A.; Bau, E.C.; Nader, H.B.; Dietrich, C.P. Distribution of sulfated glycosaminoglycans in the animal kingdom: Widespread occurrence of heparin-like compounds in invertebrates. Biochim. Biophys. Acta 2000, 1475, 287–294. [Google Scholar] [CrossRef]
  699. Holmes, B.B.; DeVos, S.L.; Kfoury, N.; Li, M.; Jacks, R.; Yanamandra, K.; Ouidja, M.O.; Brodsky, F.M.; Marasa, J.; Bagchi, D.P.; et al. Heparan sulfate proteoglycans mediate internalization and propagation of specific proteopathic seeds. Proc. Natl. Acad. Sci. USA 2013, 110, E3138–E3147. [Google Scholar] [CrossRef] [Green Version]
  700. Skaanning, L.K.; Santoro, A.; Skamris, T.; Martinsen, J.H.; D’Ursi, A.M.; Bucciarelli, S.; Vestergaard, B.; Bugge, K.; Langkilde, A.E.; Kragelund, B.B. The non-fibrillating N-terminal of alpha-synuclein binds and co-fibrillates with heparin. Biomolecules 2020, 10, 1192. [Google Scholar] [CrossRef]
  701. Hudak, A.; Kusz, E.; Domonkos, I.; Josvay, K.; Kodamullil, A.T.; Szilak, L.; Hofmann-Apitius, M.; Letoha, T. Contribution of syndecans to cellular uptake and fibrillation of alpha-synuclein and tau. Sci. Rep. 2019, 9, 16543. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  702. Lehri-Boufala, S.; Ouidja, M.O.; Barbier-Chassefiere, V.; Henault, E.; Raisman-Vozari, R.; Garrigue-Antar, L.; Papy-Garcia, D.; Morin, C. New roles of glycosaminoglycans in alpha-synuclein aggregation in a cellular model of Parkinson disease. PLoS ONE 2015, 10, e0116641. [Google Scholar] [CrossRef]
  703. Stefanova, N.; Schanda, K.; Klimaschewski, L.; Poewe, W.; Wenning, G.K.; Reindl, M. Tumor necrosis factor-alpha-induced cell death in U373 cells overexpressing alpha-synuclein. J. Neurosci. Res. 2003, 73, 334–340. [Google Scholar] [CrossRef] [PubMed]
  704. Tsuboi, K.; Grzesiak, J.J.; Bouvet, M.; Hashimoto, M.; Masliah, E.; Shults, C.W. Alpha-synuclein overexpression in oligodendrocytic cells results in impaired adhesion to fibronectin and cell death. Mol. Cell Neurosci. 2005, 29, 259–268. [Google Scholar] [CrossRef] [PubMed]
  705. Kuzdas, D.; Stemberger, S.; Gaburro, S.; Stefanova, N.; Singewald, N.; Wenning, G.K. Oligodendroglial alpha-synucleinopathy and MSA-like cardiovascular autonomic failure: Experimental evidence. Exp. Neurol. 2013, 247, 531–536. [Google Scholar] [CrossRef] [Green Version]
  706. Stefanova, N.; Reindl, M.; Neumann, M.; Kahle, P.J.; Poewe, W.; Wenning, G.K. Microglial activation mediates neurodegeneration related to oligodendroglial alpha-synucleinopathy: Implications for multiple system atrophy. Mov. Disord. 2007, 22, 2196–2203. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The role of aSyn at the presynaptic terminal. A schematic representation depicting of aSyn physiological and pathological effects at the synapse: (a) aSyn reduces the activity of tyrosine hydroxylase (TH), the enzyme responsible for catalyzing the conversion of L-Tyrosine to L-DOPA, thus impairing dopamine biosynthesis, (b) Increased levels of aSyn inhibit VMAT2, which is responsible for the uptake of monoamines (such as dopamine) into synaptic vesicles, (c) aSyn associates with synaptic vesicle membranes and regulates the SNARE-dependent vesicle fusion and neurotransmitter release, (d) Soluble aSyn interacts with the dopaminergic transporter DAT and decreases its amount on the plasma membrane, thus regulating the dopamine re-uptake from the synapse. However, aSyn aggregates trigger DAT recruitment to the plasma membrane that leads to massive entry of dopamine, (e) aSyn aggregates interact with Na+/K+-ATPase (NKA) preventing the effective pump out of Na+ ions, (f) aSyn is secreted from neuronal cells partly via associating with exosomes, (g) Extracellular aSyn interacts with neuronal receptors (i.e., LAG3) for its internalization in neurons or (h) it is up-taken via endocytosis, (i) PD-linked A30P and A53T mutant aSyn form large membrane pores through which most cations (i.e., Ca2+) can pass non-selectively, (j) Extracellular aSyn activates the voltage-gated Ca2+ channels (VGCCs), resulting in increased Ca2+ influx, (k) Monomeric aSyn enters neuronal cells via passive diffusion or direct penetration of their plasma membrane.
Figure 1. The role of aSyn at the presynaptic terminal. A schematic representation depicting of aSyn physiological and pathological effects at the synapse: (a) aSyn reduces the activity of tyrosine hydroxylase (TH), the enzyme responsible for catalyzing the conversion of L-Tyrosine to L-DOPA, thus impairing dopamine biosynthesis, (b) Increased levels of aSyn inhibit VMAT2, which is responsible for the uptake of monoamines (such as dopamine) into synaptic vesicles, (c) aSyn associates with synaptic vesicle membranes and regulates the SNARE-dependent vesicle fusion and neurotransmitter release, (d) Soluble aSyn interacts with the dopaminergic transporter DAT and decreases its amount on the plasma membrane, thus regulating the dopamine re-uptake from the synapse. However, aSyn aggregates trigger DAT recruitment to the plasma membrane that leads to massive entry of dopamine, (e) aSyn aggregates interact with Na+/K+-ATPase (NKA) preventing the effective pump out of Na+ ions, (f) aSyn is secreted from neuronal cells partly via associating with exosomes, (g) Extracellular aSyn interacts with neuronal receptors (i.e., LAG3) for its internalization in neurons or (h) it is up-taken via endocytosis, (i) PD-linked A30P and A53T mutant aSyn form large membrane pores through which most cations (i.e., Ca2+) can pass non-selectively, (j) Extracellular aSyn activates the voltage-gated Ca2+ channels (VGCCs), resulting in increased Ca2+ influx, (k) Monomeric aSyn enters neuronal cells via passive diffusion or direct penetration of their plasma membrane.
Ijms 22 04994 g001
Figure 2. The proposed intracellular effects of various aSyn conformations in neurons. A schematic representation of the aberrant interactions between the various aSyn species with intracellular organelles: (a) In the cytoplasm of neurons, aSyn monomers form oligomers that can eventually become fibrils, (b) Both unfolded and aggregated aSyn impair the function of PMR1, a Ca2+-transporting ATPase pump that regulates Ca2+ and Mn+2 levels in the Golgi apparatus (GA), resulting in elevated cytosolic Ca2+ levels, (c) Both WT and mutant A53T aSyn disrupt the vesicular transport from Endoplasmic Reticulum (ER) to Golgi (GA), (d) WT aSyn inhibits the transportation of methyltransferases from the cytoplasm to the nucleus (N), thus altering DNA methylation of the SNCA gene, (e) Inside the nucleus (N), aSyn inhibits histone acetylation via its direct binding to histones or by preventing the action of histone acetyltransferase (HAT) enzymes, thus interfering in the process of gene transcription, (f) In the ER, aSyn aggregates activate the Ca2+-ATPase SERCA, resulting in dysregulated Ca2+ homeostasis, (g) Both monomeric and oligomeric aSyn interact with Voltage-dependent anion channel 1 (VDAC1) and inositol triphosphate receptors (IP3Rs), the protein components involved in mitochondrial-associated ER membrane (MAM) and regulates the transmission of Ca2+ signals from the ER to mitochondria (M), (h) aSyn binds to TOM20, a mitochondrial import receptor subunit and inhibits normal protein import, (i) Normally, monomeric or dimeric forms of aSyn are degraded in the lysosome (L) via Chaperone Mediated Autophagy (CMA), following their interaction with LAMP2A. However, under pathological conditions, impairment of CMA has been proposed to lead to aSyn accumulation and subsequent cell toxicity, (j) Oligomeric aSyn and various misfolded proteins are cleared via macroautophagy, following the fusion of autophagosomes with the lysosome. Pathological aSyn has been shown to inhibit autophagosome maturation or their fusion with lysosomes, thus impairing autophagic flux, (k) Monomeric and oligomeric aSyn are degraded via the proteasome; however, under pathological conditions, increased levels of aSyn or even soluble aSyn oligomers may inhibit proteasomal function, leading to aSyn accumulation and the formation of insoluble aggregates.
Figure 2. The proposed intracellular effects of various aSyn conformations in neurons. A schematic representation of the aberrant interactions between the various aSyn species with intracellular organelles: (a) In the cytoplasm of neurons, aSyn monomers form oligomers that can eventually become fibrils, (b) Both unfolded and aggregated aSyn impair the function of PMR1, a Ca2+-transporting ATPase pump that regulates Ca2+ and Mn+2 levels in the Golgi apparatus (GA), resulting in elevated cytosolic Ca2+ levels, (c) Both WT and mutant A53T aSyn disrupt the vesicular transport from Endoplasmic Reticulum (ER) to Golgi (GA), (d) WT aSyn inhibits the transportation of methyltransferases from the cytoplasm to the nucleus (N), thus altering DNA methylation of the SNCA gene, (e) Inside the nucleus (N), aSyn inhibits histone acetylation via its direct binding to histones or by preventing the action of histone acetyltransferase (HAT) enzymes, thus interfering in the process of gene transcription, (f) In the ER, aSyn aggregates activate the Ca2+-ATPase SERCA, resulting in dysregulated Ca2+ homeostasis, (g) Both monomeric and oligomeric aSyn interact with Voltage-dependent anion channel 1 (VDAC1) and inositol triphosphate receptors (IP3Rs), the protein components involved in mitochondrial-associated ER membrane (MAM) and regulates the transmission of Ca2+ signals from the ER to mitochondria (M), (h) aSyn binds to TOM20, a mitochondrial import receptor subunit and inhibits normal protein import, (i) Normally, monomeric or dimeric forms of aSyn are degraded in the lysosome (L) via Chaperone Mediated Autophagy (CMA), following their interaction with LAMP2A. However, under pathological conditions, impairment of CMA has been proposed to lead to aSyn accumulation and subsequent cell toxicity, (j) Oligomeric aSyn and various misfolded proteins are cleared via macroautophagy, following the fusion of autophagosomes with the lysosome. Pathological aSyn has been shown to inhibit autophagosome maturation or their fusion with lysosomes, thus impairing autophagic flux, (k) Monomeric and oligomeric aSyn are degraded via the proteasome; however, under pathological conditions, increased levels of aSyn or even soluble aSyn oligomers may inhibit proteasomal function, leading to aSyn accumulation and the formation of insoluble aggregates.
Ijms 22 04994 g002
Figure 3. A complicated neuron-glial interrelationship. (a) Oligomeric aSyn enters oligodendrocytes (blue) via binding to the gap junction protein connexin-32 (Cx32), (b) Neuronally-derived aSyn enters oligodendrocytes via clathrin-mediated endocytosis, exosomal transportation, or via binding to Heparan Sulfate ProteoGlycans (HSPGs), (c) Inside oligodendrocytes, neuronal aSyn (red) initiates the seeding of the endogenous oligodendroglial aSyn (black) and together with the oligodendroglial-specific TPPP/p25α protein, they lead to the formation of GCIs, (d) aSyn aggregates impair the proteolytic machineries of oligodendrocytes [(proteasome and lysosome (L)], (e) Misfolded aSyn leads to mitochondrial (M) dysfunction and subsequent cell toxicity in MSA, (f) Oligodendroglial-derived exosomes containing aSyn have been shown to transfer to microglial cells (orange), (g) Extracellular aSyn stimulates TLRs and other receptors (i.e., FcγR, P2X7 etc) to activate microglial transcription factors, such as NFkB, for the production of various pro-inflammatory cytokines (IL1β, TNFa) and chemokines that induce astrocyte (green) reactivity and oligodendroglial damage, (h) Neuronally-secreted aSyn is taken up by microglial cells via clathrin-mediated endocytosis and is then driven to the lysosome (L) for degradation. In pathological conditions, though, it accumulates into aSyn insoluble aggregates, (i) Free or exosome-associated aSyn released by neurons is transmitted to astrocytes via endocytosis and enters the lysosome (L) for its clearance, (j) In disease, aSyn aggregates are formed that lead to lysosomal (L) impairment and mitochondria (M) dysfunction, (k) In astrocytes, aSyn triggers the opening of channels (i.e., Cx43 and Panx1), leading to dysregulation in Ca2+ homeostasis and altered mitochondrial morphology, (l) Microglia-secreted cytokines and chemokines activate astrocytes to further produce pro-inflammatory signaling molecules and enhance neurotoxicity.
Figure 3. A complicated neuron-glial interrelationship. (a) Oligomeric aSyn enters oligodendrocytes (blue) via binding to the gap junction protein connexin-32 (Cx32), (b) Neuronally-derived aSyn enters oligodendrocytes via clathrin-mediated endocytosis, exosomal transportation, or via binding to Heparan Sulfate ProteoGlycans (HSPGs), (c) Inside oligodendrocytes, neuronal aSyn (red) initiates the seeding of the endogenous oligodendroglial aSyn (black) and together with the oligodendroglial-specific TPPP/p25α protein, they lead to the formation of GCIs, (d) aSyn aggregates impair the proteolytic machineries of oligodendrocytes [(proteasome and lysosome (L)], (e) Misfolded aSyn leads to mitochondrial (M) dysfunction and subsequent cell toxicity in MSA, (f) Oligodendroglial-derived exosomes containing aSyn have been shown to transfer to microglial cells (orange), (g) Extracellular aSyn stimulates TLRs and other receptors (i.e., FcγR, P2X7 etc) to activate microglial transcription factors, such as NFkB, for the production of various pro-inflammatory cytokines (IL1β, TNFa) and chemokines that induce astrocyte (green) reactivity and oligodendroglial damage, (h) Neuronally-secreted aSyn is taken up by microglial cells via clathrin-mediated endocytosis and is then driven to the lysosome (L) for degradation. In pathological conditions, though, it accumulates into aSyn insoluble aggregates, (i) Free or exosome-associated aSyn released by neurons is transmitted to astrocytes via endocytosis and enters the lysosome (L) for its clearance, (j) In disease, aSyn aggregates are formed that lead to lysosomal (L) impairment and mitochondria (M) dysfunction, (k) In astrocytes, aSyn triggers the opening of channels (i.e., Cx43 and Panx1), leading to dysregulation in Ca2+ homeostasis and altered mitochondrial morphology, (l) Microglia-secreted cytokines and chemokines activate astrocytes to further produce pro-inflammatory signaling molecules and enhance neurotoxicity.
Ijms 22 04994 g003
Figure 4. Crosstalk between astrocytes and microglia in a-synucleinopathies. Monomeric or oligomeric aSyn released by neurons is taken-up by astrocytes (purple) and microglia (red), which are then activated and secrete various cytokines or chemokines. (a) According to the prevailing hypothesis, the released pro-inflammatory molecules by microglia trigger the activation of astrocytes (red arrow), leading to excessive inflammation and neurotoxicity. (b) Conversely, aSyn can directly activate astrocytes to secrete pro-inflammatory cytokines or chemokines that recruit and activate microglia (purple arrow) resulting in excessive neuroinflammation.
Figure 4. Crosstalk between astrocytes and microglia in a-synucleinopathies. Monomeric or oligomeric aSyn released by neurons is taken-up by astrocytes (purple) and microglia (red), which are then activated and secrete various cytokines or chemokines. (a) According to the prevailing hypothesis, the released pro-inflammatory molecules by microglia trigger the activation of astrocytes (red arrow), leading to excessive inflammation and neurotoxicity. (b) Conversely, aSyn can directly activate astrocytes to secrete pro-inflammatory cytokines or chemokines that recruit and activate microglia (purple arrow) resulting in excessive neuroinflammation.
Ijms 22 04994 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mavroeidi, P.; Xilouri, M. Neurons and Glia Interplay in α-Synucleinopathies. Int. J. Mol. Sci. 2021, 22, 4994. https://doi.org/10.3390/ijms22094994

AMA Style

Mavroeidi P, Xilouri M. Neurons and Glia Interplay in α-Synucleinopathies. International Journal of Molecular Sciences. 2021; 22(9):4994. https://doi.org/10.3390/ijms22094994

Chicago/Turabian Style

Mavroeidi, Panagiota, and Maria Xilouri. 2021. "Neurons and Glia Interplay in α-Synucleinopathies" International Journal of Molecular Sciences 22, no. 9: 4994. https://doi.org/10.3390/ijms22094994

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