Skip to main content
Neural Plasticity logoLink to Neural Plasticity
. 2017 Jan 3;2017:5012129. doi: 10.1155/2017/5012129

The Contribution of α-Synuclein Spreading to Parkinson's Disease Synaptopathy

Francesca Longhena 1, Gaia Faustini 1, Cristina Missale 1, Marina Pizzi 1,2, PierFranco Spano 2, Arianna Bellucci 1,*
PMCID: PMC5241463  PMID: 28133550

Abstract

Synaptopathies are diseases with synapse defects as shared pathogenic features, encompassing neurodegenerative disorders such as Parkinson's disease (PD). In sporadic PD, the most common age-related neurodegenerative movement disorder, nigrostriatal dopaminergic deficits are responsible for the onset of motor symptoms that have been related to α-synuclein deposition at synaptic sites. Indeed, α-synuclein accumulation can impair synaptic dopamine release and induces the death of nigrostriatal neurons. While in physiological conditions the protein can interact with and modulate synaptic vesicle proteins and membranes, numerous experimental evidences have confirmed that its pathological aggregation can compromise correct neuronal functioning. In addition, recent findings indicate that α-synuclein pathology spreads into the brain and can affect the peripheral autonomic and somatic nervous system. Indeed, monomeric, oligomeric, and fibrillary α-synuclein can move from cell to cell and can trigger the aggregation of the endogenous protein in recipient neurons. This novel “prion-like” behavior could further contribute to synaptic failure in PD and other synucleinopathies. This review describes the major findings supporting the occurrence of α-synuclein pathology propagation in PD and discusses how this phenomenon could induce or contribute to synaptic injury and degeneration.

1. Introduction

Pathological accumulation of α-synuclein in the brain is a typical neuropathological hallmark of Parkinson's disease (PD), a complex neurodegenerative disorder characterized by motor disability that derives from the neurodegeneration of nigrostriatal dopaminergic neurons. The presence of Lewy bodies (LB) and Lewy neurites (LN), proteinaceous inclusions whose main constituent is fibrillary-aggregated α-synuclein, is a defining neuropathological alteration observed in the brain of affected patients. In the last few years, it has become evident that PD may be considered as a synaptopathy [1, 2]. Indeed, striatal dopaminergic terminal loss appears to precede neurodegeneration in the substantia nigra [3, 4] and the deposition of α-synuclein, which is considered as a causative factor for the onset of the disorder, mainly affects synaptic terminals in its early stages [5, 6]. Nonetheless, the exact molecular mechanisms that determine the selective vulnerability of nigrostriatal synapses to α-synuclein deposition are still enigmatic. One of the reasons for the reduced resilience of dopaminergic terminals to α-synuclein accumulation may be the fact that dopamine neurons of the substantia nigra show relatively large axonal lengths with 10–100 times the number of synapses of neighboring neuronal cells [7]. This implies that these cells require an elevated energy demand to efficiently sustain the trafficking of organelles and vesicles to allow proper synapse functioning [8]. Hence, they might result particularly susceptible even to subtle homeostatic changes at synaptic sites, which constitute their major source of energy consumption.

Compelling evidence indicates that α-synuclein can spread from diseased to healthy cells, thus contributing to disease worsening [9]. Indeed, graft-derived dopamine neurons can develop LB pathology several years after transplantation [1012], an event that can be responsible for the loss of beneficial effects of cell therapy along with time. In addition, both neurons and astrocytes have been found to internalize disease-associated α-synuclein in the postmortem brain of subjects with Lewy pathology, suggesting prion-like spread of α-synuclein by uptake from surrounding structures [13].

It is worth noting that the progression of PD symptoms seems to correlate with the topographical brain propagation of α-synuclein pathology between anatomically interconnected brain areas [14, 15], suggesting the occurrence of trans-synaptic spreading of pathology. This is not surprising when considering that α-synuclein is a synaptic-vesicle associated protein [16]. Studies on experimental models of PD have confirmed the occurrence of trans-synaptic transmission of pathological α-synuclein in vivo [1720]. In light of evidences supporting host-to-graft α-synuclein passage, as well as the ability of the protein to propagate protein misfolding in recipient cells, a “prion-like” hypothesis of PD has been postulated [2129]. Indeed, prions are transmissible misfolded conformers of the prion protein, PrP, which seed further generation of infectious proteins [30]. However, the mechanisms mediating α-synuclein release, uptake, and behavior in recipient cells deserve further investigation to definitely assert that α-synuclein behaves like a prion. For instance, not all the postmortem analyses on PD cases seem to confirm the caudo/rostral spread of α-synuclein pathology throughout the brain described by Braak [14, 15]. Moreover, the ability to seed aggregation in recipient cells has been found to be dependent on high concentration of aggregates in the face of the need of low amount of aggregates to induce cytotoxicity [31]. These evidences pose into question both the trans-synaptic spreading hypothesis and the prion-like behavior of α-synuclein. Nonetheless, cell-to-cell transmission of α-synuclein can occur with Ca2+-dependent exosome-mediated release [32] or nonclassical exocytosis [33] as the most plausible candidate mechanisms for α-synuclein ejection. Finally, what could be the transmissible form of the protein that can mediate toxicity and/or function as seed for “misfolding” propagation in recipient cells between monomeric, oligomeric, and fibrillary α-synuclein still needs to be clarified. In this review, we describe the features of α-synuclein pathology spreading in PD and discuss how this could contribute to synaptic damage.

2. Alpha-Synuclein Function at the Dopamine Synapse

Αlpha-synuclein was first identified in the synaptic vesicles and nuclei of the electric organ of Torpedo Californica [34]. The protein is highly expressed in presynaptic terminals of the brain and peripheral nervous system, where it associates with the synaptic vesicle apparatus [3537]. The expression of α-synuclein is elevated within the synapses of nigral dopamine neurons [38]. There, the protein can modulate synaptic dopamine release by directly modulating the dopamine transporter (DAT), synapsin III, the small GTP-binding protein Rab3A, and the soluble N-ethylmaleimide sensitive fusion attachment protein receptor (SNARE) protein member vesicle associated membrane protein-2 (VAMP-2) [6]. Indeed, it catalyzes the entry of VAMP-2 into the SNARE complex [39] and enhances DAT localization at the plasma membrane [40], which consistently is impaired by α-synuclein aggregation [41, 42]. The toxicity exerted by α-synuclein deposition in dopaminergic neurons is rescued by Rab3A expression, suggesting that this protein is relevant for its normal function [43]. Conversely, α-synuclein membrane association is regulated by the Rab3A recycling machinery and presynaptic activity [44]. Finally, synapsin III distribution and expression in dopaminergic neurons is regulated by α-synuclein, while synapsin III gene silencing inhibits α-synuclein aggregation [4].

In the synapse, likewise in the other neuronal compartments, α-synuclein exists in a dynamic equilibrium between a soluble state and a membrane-bound state, with its affinity for synaptic vesicles being higher than that for cell membranes [4547] (Figure 1(a)). However, when interacting with cell membranes, the protein presents higher affinity for lipid rafts [48]. The interaction between α-synuclein and lipid membranes is also relevant for the protein to exert its functions [49]. This, notwithstanding, α-synuclein presents elevated structural plasticity and its effective conformation in soluble and membrane-bound state is matter of debate. While some evidences support the existence of a stable cytosolic tetrameric form of the protein, other studies have shown that it can be found as a disordered monomer in the central nervous system (CNS) and other mammalian cells or that these forms both coexist in a dynamic equilibrium [50, 51]. More recently, the existence of a stable dimer has been suggested by computational evidences [52]. As for the aminoacidic sites involved in the interaction with membranes, numerous studies have reported a key role for the ones located at the N-terminal portion of α-synuclein [53]. Jiang and coauthors [54], by coupling neutron reflectometry (NR) and fluorescence spectroscopy analysis, have found that the N- and C-terminal regions near positions 4 and 94 are anchored to the membrane, while the putative linker spanning residue 39 samples multiple conformations, which are sensitive to the chemical nature of the membrane surface. The mechanism of α-synuclein binding to lipid membranes has been found to be primarily dependent on the surface charge density of the lipid bilayer and the phase state of the lipids, with α-synuclein possessing lipid ordering effect and thermally stabilizing vesicles [55]. These findings suggest that this process might be tunable by environmental changes.

Figure 1.

Figure 1

Monomeric, oligomeric, and fibrillary α-synuclein at the synaptic terminal. (a) Monomeric α-synuclein modulates synaptic function by controlling synaptic vesicle release. This form of the protein can be released in association with exosomes, activates microglial cells, and can be internalized at postsynaptic sites. (b) Oligomeric α-synuclein formation is enhanced by interaction of monomeric protein with DA. Alpha-synuclein oligomers can form a stable adduct with the toxic dopamine metabolite DOPAL. Oligomers can be released in association with extracellular vesicles and then activate microglia. Alpha-synuclein oligomers can disrupt synaptic vesicles membranes as well as presynaptic and postsynaptic membranes. Exogenous α-synuclein oligomers can damage lipid rafts and affect LTP by activating NMDA receptors. Intracellular α-synuclein oligomers with endogenous or exogenous origin impair mitochondrial functions and cytoskeletal architecture. (c) Fibrillary-aggregated α-synuclein alters synaptic vesicle release by clustering synaptic vesicles and by perforating plasma membrane. Extracellular fibrils deriving from degenerating neurons in the PD brain can activate microglial cells and actively contribute to alpha-synuclein pathology spreading. The formation of endogenous α-synuclein fibrils can reduce seeding activity and toxicity although exogenous α-synuclein fibrils function as a seed for the aggregation of endogenous α-synuclein in recipient cells.

Collectively, these evidences support that subtle changes in α-synuclein structural folding, likewise the formation of oligomers or insoluble aggregates, can severely compromise the activity of dopaminergic neurons, with the adjunct of cell-to-cell transmission likely worsening and perpetrating the related injury.

3. Central and Peripheral Localization of Neuronal α-Synuclein Pathology in PD: Trans-Synaptic or Systemic Spreading?

The presence of different forms of α-synuclein in cerebrospinal fluid (CSF), blood plasma [5658], and saliva [59] coupled to the discovery of α-synuclein pathology in embryonic nigral transplants of PD patients [10, 11, 60] suggested that α-synuclein can move from cell to cell and can initiate pathology in recipient neurons. It is well established that α-synuclein can be secreted [61, 62], a process that is thought to be mainly mediated by exosomal vesicles [63, 64].

Compelling evidence indicates that in PD α-synuclein pathology is not confined within the brain. LB-like aggregates of the protein have been found in the dorsal root ganglia, as well as in several tracts of the peripheral nervous system such as gastrointestinal innervation, motor nerves innervating the pharyngeal muscles, cranial and spinal nerves, skin nerves, and olfactory epithelium [8, 6570]. The presence of these peripheral aggregates has been proposed to associate with typical PD premotor symptoms [71] and could contribute to disease onset in the brain, as supported by evidences indicating that vagotomy diminishes the risk to develop the disorder [20]. As hypothesized by Braak and colleagues [14], caudo-rostral spreading of α-synuclein has been found to occur through vagal connections in experimental models of PD [17]. Resection of the vagal autonomic projection impedes the peripheral-to-CNS diffusion of pathological α-synuclein and the onset of PD-like motor phenotype in the chronic intragastric rotenone model of PD [19], supporting a causative role for protein spreading in the onset of the disorder. Vagotomy eliminates most, but not all, α-synuclein-positive neurites in the plexus, thus providing a candidate pathway for the retrograde transport of putative PD pathogens or toxins from the enteric nervous system to the central nervous system [67]. However, even vagal nerve impairment per se has been found to induce dopamine functional damage, therefore suggesting that the vagal degeneration occurring in the PD brain [72, 73] might be pivotally involved in PD pathogenesis independently from α-synuclein pathological spreading. Studies in rodents have shown that vagal afferent endings in the myenteric plexus and the gastrointestinal smooth muscle do not express α-synuclein, whereas virtually all vagal preganglionic projections to the gut show α-synuclein immunopositivity both in axons and in terminal varicosities in apposition with myenteric neurons.

However, some other studies raise some concerns about its validity as α-synuclein accumulation within the peripheral nervous system can occur also in neurologically intact subjects with aging [74]. Not by chance, PD and aging have been proposed to be a unique entity and the disease has been postulated to manifest in all subjects whether they could live long enough [75]. In addition, typical manifestations of aging such as frailty that interestingly has been found to associate with brain neuropathological accumulation of LB and nigral neuronal loss [76, 77] also characterize PD [78]. Other studies, failing to detect either the typical pattern of topographical distribution of α-synuclein pathology in the postmortem brain of affected patients [79, 80] or the development of pathology spreading following preformed fibril-inoculation in vivo, cast further confusion over the prion-like hypothesis.

Recent studies indicate that α-synuclein oligomers are increased in red blood cells and CSF of PD subjects and could serve as biomarkers of disease [81, 82]. The levels of α-synuclein are also increased in peripheral lymphocytes [83] as well as in plasma and CNS-derived exosomes of affected individuals [84]. The systemic spreading of the protein could thus also involve exosomal-mediated release and biological fluids. Indeed exosomes are small membrane vesicles which result from exocytosis of multivesicular bodies. They function as mediators of intercellular communication, as they transfer specific proteins, lipids, microRNAs, and DNA between cells. Because of their small size, exosomes can move from the site of discharge by diffusion and reach several biological fluids, such as blood, CSF, urine, and synovial fluid [85]. Consistently, plasma exosomal α-synuclein is likely CNS-derived and increased in PD [84] and CSF exosomes have been found to induce α-synuclein aggregate formation in recipient healthy cells [86]. This suggests that the circulatory system could also mediate α-synuclein systemic spreading, with the choroid plexus being likely involved in the passage of alpha-synuclein from the blood to the brain and vice versa. Notably, the presence of increased α-synuclein immunoreactivity at this site has been described in PD [13]. However, it still remains to be determined whether a link effectively exists between peripheral and brain α-synuclein deposition. In the case that they might constitute a unique phenomenological entity, defining what might come first between central and peripheral deposition could be crucial for determining the causes of PD.

4. Toxicity of α-Synuclein Oligomers and Fibrils: A Still Unresolved Dichotomy

Aggregation of α-synuclein is a critical step in the etiology of PD, with prefibrillar oligomers of the protein that might constitute the direct precursors of fibrils being involved in neurodegenerative process [87, 88]. Even if the injection of fibrils into the rat brain is found to be more toxic than that of oligomers and ribbons, as it induces neurodegeneration and motor impairment, all these species can self-amplify in vivo and lead to PD/multiple system atrophy-like alterations in the injected animals [89]. This suggests that strain-specific prion-like infectivity and symptomatology characterize different α-synuclein conformers, whose biochemical nature is still unknown. Therefore, asserting which, between fibrils or oligomers, are the most toxic species is still difficult at present. Nonetheless, the analysis of both oligomers' and fibrils' structure could help to elucidate this conundrum. In addition, data supporting that monomeric, oligomeric, and fibrillary α-synuclein can activate microglial cells [9093] suggest that all these forms of the protein could affect neuronal homeostasis by modulating microglia function that could be either protective or detrimental in PD [94].

As for oligomers, those formed by a peptide derived from residues 36–55 of α-synuclein were recently characterized by X-ray crystallography [95]. The authors showed that this specific peptide is able to adopt a β-hairpin structure, which assembles in a hierarchical fashion. Three β-hairpins then assemble to form a triangular trimer and three copies of the triangular trimer then assemble to form a basket-shaped nonamer that couple to form an octadecamer. Following molecular modeling analysis, these authors also proposed that full-length α-synuclein might be able to assemble in this fashion. Circular dichroism spectroscopy demonstrated that the peptide 36–55 interacts with anionic lipid bilayer membranes, like oligomers of full-length α-synuclein, and is found to be toxic in neuronal cell models. Other cryoelectron microscopy studies have shown that oligomers isolated during fibril formation possess a hollow cylindrical architecture with similarities to certain types of amyloid fibril [96].

The formation of α-synuclein dimers has been described to initiate aggregation and neurotoxicity [97]. Computer modeling and cell-based studies have revealed that upon interaction with plasma membranes α-synuclein rapidly penetrates them, changing its conformation from α-helical toward a coiled structure. This penetration facilitates the incorporation of additional α-synuclein monomers to the complex and subsequent displacement of phospholipids and formation of oligomers in the membrane [98]. Consistently, α-synuclein oligomers neurotoxicity in vivo has been found to be mediated by membrane disruption [99].

Chen and coauthors described that stable toxic α-synuclein oligomeric species with a hollow cylindrical architecture with similarities to certain types of amyloid fibril can be trapped during fibril formation [96]. Their study showed that the β-sheet geometry acquired in the early stages of the self-assembly process plays a key role in dictating the kinetic stability and the pathological nature of individual oligomeric species. Spectroscopy studies have also shown that whereas fibrils adopt a parallel arrangement oligomers adopt an antiparallel β-sheet structure [100], thus suggesting that differences in the toxicity of these species might rely on their diverse conformations. The neurotoxicity of oligomers has been demonstrated in different experimental conditions. Alpha-synuclein overexpression in neuroblastoma cells causes the formation of α-synuclein oligomeric species, whose presence is associated with mitochondrial fragmentation and autophagic-lysosomal pathway activation in live cells [101]. The accumulation of oligomeric and fibrillar forms of α-synuclein has a negative impact on mitochondria function by inhibiting mitochondria complexes IV and V in the striatum [102], reducing mitochondrial Ca2+ release and NADH oxidation [103]. Permeabilization of mitochondria membranes can also be induced by α-synuclein oligomers [104107]. The accumulation of toxic α-synuclein oligomers in the endoplasmic reticulum (ER) has been described as a feature of α-synucleinopathies [108, 109]. Recently, α-synuclein oligomers have been found to interact with metal ions to induce oxidative stress and neuronal death in PD [110]. Interestingly, the oligomer-induced reactive oxygen species (ROS) production was independent of several known cellular enzymatic sources and relied solely on the presence of free metal ions.

Studies on human α-synuclein transgenic mice have shown that the accumulation of oligomers mainly occurs at synaptic sites and is crucial for the induction of synaptic and neuronal degeneration [111]. Consistently, the formation of α-synuclein soluble oligomers can reduce neuronal excitability of neocortical pyramidal cells, suggesting that it could impact on network activity [112]. Oligomers have been found to affect synaptic function by changing lipid raft composition and increasing N-methyl-D-aspartate (NMDA) receptor activation at postsynaptic membranes [113, 114].

Alpha-synuclein oligomers could thus easily impinge on synaptic resilience by disrupting membranes, inducing mitochondrial depolarization, altering cytoskeleton dynamics, impairing protein clearance pathways, and enhancing oxidative stress [115] (Figure 1(b)). In addition, the preferential degeneration of dopaminergic neurons in PD [116, 117] led to the hypothesis that dopamine may play an important role in α-synuclein oligomerization [118]. The oxide forms of dopamine, generated by oxidative stress, accelerate formation of α-synuclein oligomers as an endogenous protofibril stabilizer, demonstrating the connection between dopamine and α-synuclein oligomer formation [119, 120]. The toxic dopamine metabolite 3,4-dihydroxyphenylacetaldehyde (DOPAL) can compromise the membrane-binding affinity of α-synuclein to synaptic vesicles, as well as its fibrillation, by forming an adduct with the protein, thus reducing its functional ability to modulate synaptic vesicles [121].

Fibrils exert their toxic actions by activating other detrimental processes in neuronal cells. This may be related to the different conformation of oligomers and fibrils. Consistently, Curtain and coauthors [122], by using small angle X-ray scattering and ensemble optimization modeling studies, were able to demonstrate that α-synuclein oligomers and fibrils originate in two distinct conformer pools, with E53T and E45K mutations enhancing the tendency to form fibrils, while the A30P conferring propensity toward oligomer assembly. This is in line with previous data reporting that A30P mutant α-synuclein forms different fibril structures [123] and that mutations in the KTKE(Q)GV imperfect amino acid repeats in the N-terminal part of the protein can also affect its tendency toward fibril formation [124]. Mutant α-synuclein preferentially shifts from monomer to fibrils, thus suggesting that the formation of these species plays a crucial role in the pathophysiology of early-onset PD [125, 126]. Whether the lipid bound α-helical form or the unfolded state initiates protein aggregation remains to be determined.

X-ray and electron diffraction of the α-synuclein aggregates showed that the fibrils consist of a β-sheet structure in which the β-strands run perpendicularly to the long fiber axis [127]. A flexible break close to residues 52–55 has been found to be relevant for fibril formation [128], while the negative charges and aromatic residues at the C-terminal region play a modulatory role on fibrillation [129, 130]. In addition, molecular dynamics simulations allowed observing that residues 36−55 of the nonamyloid component (NAC) domain, in the central region of the protein, are important for the formation of β-hairpin and that the point mutations stabilize this β-hairpin that represented the first step of α-synuclein aggregation [95].

Αlpha-synuclein fibrils can exert toxicity by disrupting the normal distribution of membrane proteins and synaptic vesicles [6], perforating plasma membranes [131, 132], triggering cellular responses [133], and causing death in cell systems [134] (Figure 1(c)). However, the stabilization of fibril clusters can prevent fragmentation and reduces seeding activity and toxicity [135]. In addition, α-synuclein structure is strictly dependent on membrane interactions that can both accelerate [136, 137] and inhibit [136, 138, 139] fibril formation.

5. Mechanisms of α-Synuclein Release

Since 2003, Braak and colleagues had speculated that, in idiopathic PD, LB pathology could spread from the enteric nervous system or the olfactory bulb to precise brain regions during the progression of the disease [14], thus suggesting the occurrence of cell-to-cell α-synuclein transmission. The existence of this phenomenon was confirmed thanks to postmortem studies carried out on the brain of PD patients who had received transplants of mesencephalic neurons 14 years prior to the demise. Indeed, the analysis of autopsy samples showed that grafted neurons had developed LB pathology [10, 60]. Host-to-graft transmission was then corroborated in different experimental models in vivo [140142].

The first evidence in favor of α-synuclein release was provided by Dixon and colleagues that demonstrated that, in yeasts, the protein can be delivered to the extracellular space through the classical ER-to-Golgi secretory pathway [143]. However, these findings were not confirmed in SH-SY5Y cells, where α-synuclein has been observed to be mediated by ER/Golgi-independent unconventional exocytosis pathway [33, 62] and in a Ca2+-dependent manner by exosomes [63], respectively. Protein overexpression, misfolding, and posttranslational modifications can promote α-synuclein accumulation and release, caused by impaired intracellular degradation. Several cellular mechanisms have been found to mediate α-synuclein degradation, including the ubiquitin-proteasome system (UPS) and lysosome-mediated digestion [144]. In PD, dysfunction of the UPS causes the accumulation of misfolded α-synuclein and is thought to be highly implicated in the pathogenesis of PD [145149]. However, ubiquitination of α-synuclein in LB has been found to constitute a pathological event not associated with impairment of proteasomal function [150]. Alpha-synuclein can also be cleared by the autophagy-lysosome pathway (ALP). Indeed, Cuervo [151] showed that wild type α-synuclein is translocated into lysosome for degradation. Conversely, aggregated α-synuclein can cause a malfunction in lysosomal degradation pathway [152]. Remarkably, also aging may represent another important factor related to the dysfunction of protein control systems [145, 151, 153155] causing a vicious circle that exacerbate the accumulation of toxic forms of α-synuclein.

The oxidized forms of α-synuclein are preferentially secreted [62] and studies in induced pluripotent stem cell-derived neurons harboring α-synuclein gene locus triplications showed that increased intracellular levels of the protein foster its release [156]. However, the amount of released α-synuclein deriving from injured neurons is likely very low as neuronal damage and neurodegeneration do not exacerbate its propagation [18].

At the subcellular level, it has been observed that, in neurons, α-synuclein can be found in the three endosomal compartments and translocate across them (early, late, and recycling endosomes) [32]. This pathway could be involved in the process of protein release [32]. Moreover, treatments that could disturb the homeostasis of the endosomal system alter α-synuclein spreading [63, 157]. From the early endosomes, a portion of the protein can be translocated to the recycling endosomal compartment, which can fuse with the plasma membrane to allow the release of soluble forms of α-synuclein. This process seems to be mediated by the interaction between α-synuclein and Rab GTPases [158]. The portion of protein that remains in the endosomal compartment is targeted to multivesicular bodies (MVBs), which usually fuse with lysosomes for protein degradation. In diseased neurons, where the accumulation of toxic α-synuclein aggregates could perturb the activity of protein clearance mechanisms, MBVs can release the intraluminal vesicles in the extracellular space as exosomes [159]. Remarkably, numerous findings [63, 157, 160, 161] support that α-synuclein can be released in association with exosomes, small vesicles (40–100 nm in diameter) originating from the endosomal compartment. Exosomes can be secreted and interact with cell membranes in a cell type-dependent manner and recipient cells can internalize them through different endocytic mechanisms [162]. In several neurodegenerative diseases, exosomes seem to play the role of “garbage” carrier, acting as an alternative pathway of protein elimination when intracellular mechanisms are engulfed [163, 164]. The process through which α-synuclein is targeted to the endosomal compartment is still matter of study, with a recent work suggesting posttranslational modifications of the protein being involved. In particular, the conjugation of Small Ubiquitin like Modifier (SUMO) also defined as SUMOylation of α-synuclein triggers its internalization in exosomes [160]. To date, SUMOylation serves to regulate the solubility of aggregation-prone proteins [165, 166], as well as a ubiquitin-independent endosomal sorting complex required for transport (ESCRT) sorting signal, regulating the extracellular vesicle release of α-synuclein [160]. Interestingly, ubiquitinated forms of α-synuclein have been found to be present in LB [167, 168] and the ubiquitination of the protein has been found to increase its aggregation propensity and neurotoxicity in vitro [169, 170]. Taken together, these findings suggest that α-synuclein ubiquitination and SUMOylation may be coexisting phenomena whose equilibrium regulates the folding state and the localization of the protein. In particular, the fact that SUMOylation facilitates the exosomal-mediated release of α-synuclein [160] supports the hypothesis that exosomes might behave as possible “way out” to eliminate exceeding amount of the protein that cannot be degraded by conventional mechanisms such as proteasome or lysosome digestion. Further studies are needed to corroborate the role of exosomes in toxic α-synuclein spreading and in the progression of PD. Nonetheless, exosomes and their cell-to-cell transmission mechanisms could represent novel intriguing therapeutic targets to lessen or block the evolution of α-synuclein pathology.

6. Uptake of Aberrant α-Synuclein from Recipient Cells: Consequences on Synaptic Functions

In the brain, neurons can show differential vulnerability to α-synuclein accumulation. Indeed, even though it has been reported that in PD subjects LB pathology first affects the vagal nuclei, locus coeruleus, and olfactory bulbs, the motor symptoms appear when α-synuclein deposition reaches nigrostriatal dopamine neurons. In particular, since the striatal synaptic accumulation of the protein is much higher than its deposition in the cell bodies of the substantia nigra [171], it may be hypothesized that mechanisms promoting its synaptic translocation could be enhanced in the early phases of disease. When taking into account the trans-synaptic spreading hypothesis by Braak, it is feasible that nigrostriatal dopamine neurons might interpret the uptake of α-synuclein pathological seeds as a signal that prompts them to further promote α-synuclein trans-synaptic transfer as a response. In line with this idea, the injection of synthetic fibrils and LB extracts from PD subjects in the mouse brain has been found to induce a rapid and progressive synucleinopathy between anatomically interconnected brain regions [9, 164, 172174]. Nonetheless, other studies have reported considerable difficulties in inducing a widespread induction of α-synuclein pathology following intracerebral administration of amyloidogenic forms of the protein in mice [175]. Furthermore, several mechanisms, such as neuroinflammation, have been found to act synergistically or independently to promote the spread of pathology following fibrillary amyloidogenic and nonamyloidogenic α-synuclein [175]. This evidence calls into question that extracellular α-synuclein can catalyze aggregation and spread of intracellular protein only through a nucleation dependent conformational templating mechanism. Worthy of note, besides the fact that fragmented amyloid-like aggregates of short α-synuclein fibrils can function as seeds that trigger prion-like conversion [176], the transmission of mature fibrils between cells might be difficult in light of their dimensions and stability. It is easier to speculate that they might derive from degenerating neurons in the parenchyma. On this line, brain propagation of α-synuclein has been found to involve nonfibrillar protein species and to be enhanced in α-synuclein null mice [177]. In addition, PD-causing α-synuclein missense mutations shift native tetramers to monomers as a mechanism for disease initiation [51]. Extracellular oligomeric species that are also transmissible through exosomal vesicles [178] are highly abundant in the PD brain [115] and can promote α-synuclein aggregation in recipient cells [179]. Transmitted electron microscopy studies in the postmortem human brain of subjects affected by synucleinopathies, reporting the presence of oligomeric α-synuclein within the early-endosomal compartment of neuronal cells, are also in line with the idea that oligomers might be the transmissible species [13]. However, what could be the cause of their increased accumulation and how they trigger aggregation of endogenous α-synuclein in recipient cells still needs elucidation. The release of oligomers might serve to eliminate exceeding levels of the protein, or it could depend on the activation of plastic structural adaptive mechanisms in neurons as the protein is involved in synaptic plasticity [180, 181]. Otherwise, protein oligomers could behave as a transmissible neuronal messenger between neighboring neurons. Remarkably, the higher stability of α-synuclein oligomers and fibrils renders these species more suitable to be secreted in the extracellular space when compared to monomeric protein. Indeed, the conformation of this latter could be modulated even by subtle homeostatic changes in the microenvironment.

Alpha-synuclein uptake could very well contribute to synaptic damage. Indeed, the protein could accumulate at synapses by altering the function of endogenous proteins and engulfing the retrograde transport from the terminals that receive α-synuclein to the cell bodies. Alternatively, recipient cells might collapse as they fail to degrade internalized α-synuclein efficiently. It is feasible that the accumulation of extracellular α-synuclein in dopaminergic synaptic terminals can easily initiate synaptic failure given the relevance of the protein in the regulation of dopamine release [182]. Folding and misfolding of endogenous α-synuclein can be modulated by exogenous pathological α-synuclein forms and then affect, or be affected by, their interaction with lipid membranes [183]. Alpha-synuclein binding to lipid membranes can be either detrimental or protective to neuronal cells [184]. For instance, A30P, E46K, and A53T disease variants of α-synuclein show increased lipid binding affinity [185] although they have distinct membrane permeabilization properties [186] and can thus differentially affect membrane structure [187]. In line with this idea, exogenous α-synuclein has been found to induce lipid raft fragmentation thus leading to both pre- and postsynaptic alterations [114] and α-synuclein oligomers can impair long term potentiation (LTP) and impair synaptic transmission [113].

Finally, a recent study showed that different α-synuclein conformers can cross the blood brain barrier and distribute to the CNS after intravenous injection [89]. This novel evidence suggests that the diffusion of α-synuclein pathology might also be mediated by mechanisms other than the simple trans-synaptic spreading of the protein among interconnected brain regions.

Looking at the mechanisms of α-synuclein uptake, while the soluble monomeric protein can cross the plasma membrane or can be captured by Rab5a-dependent [188] and dynamin-dependent endocytosis [140], high order assembly can enter into recipient cells by using different endocytic pathways [169]. Among the possible mechanisms of uptake, macropinocytosis has also been explored [189]. Other authors reported that extracellular α-synuclein uptake by microglial cells is mediated by the GM1 ganglioside as well as by hitherto-unknown protein receptors in clathrin-, caveolae-, and dynamin-independent, but lipid raft-dependent processes [90]. However, the caveolae-specific protein caveolin-1 has been found to interact with and to mediate α-synuclein toxicity in neuroblastoma cells [190] thus suggesting that the possibility that the uptake of the protein might be mediated by caveolae-mediated endocytosis at least in neuronal cells deserves further investigation. More recently, mesenchymal stem cells were identified as blockers of the clathrin-mediated endocytosis of extracellular α-synuclein, a process that is controlled by the interaction with NMDA receptor [191]. In addition, the interaction between preformed α-synuclein fibrils and immune receptor lymphocyte activation gene 3 (LAG3) has been found to be essential for initiating the transneuronal propagation of α-synuclein [192]. This study clearly opens new avenues for PD therapy as LAG3 antibodies are already being tested as cancer treatments [193] and suggested that LAG3 might mediate both immune system activation and systemic spreading of pathological fibrillary α-synuclein species.

The fact that α-synuclein can be released in association with exosomes and extracellular vesicles (EV) strongly suggests that these might constitute the primary vehicles of cell-to-cell transmission of the protein, preserving it from degradation by extracellular enzymes and facilitating its correct targeting toward recipient cells. Indeed, given the high structural instability of the protein and its small size, it is quite unlikely that α-synuclein could easily survive in the brain parenchyma environment in a free and soluble form.

Collectively, these evidences strongly support that α-synuclein spreading could very well contribute to synaptic impairment in PD although the biological factors determining the selective vulnerability of nigrostriatal neurons, even on top of the “prion-like” hypothesis, are not yet clear.

7. Concluding Remarks

At present, the spreading hypothesis of α-synuclein pathology is still matter of debate. Indeed, although a plethora of research studies seem to confirm that α-synuclein can diffuse throughout the nervous system, is transmitted from cell to cell, and could induce toxicity and function as a seed for the aggregation of endogenous protein, other evidences seem to confute these findings (Table 1). This, notwithstanding, data on the transmissibility of the protein as well as on the misfolding-inducing ability of α-synuclein oligomers and fibrils suggested that these features might easily perturb synaptic homeostasis, especially in dopamine neurons. It could be feasible that α-synuclein release, coupled to fibrillary insoluble inclusion formation, could deprive dopamine synaptic terminals from the modulatory action of the protein. In parallel, the exosome-mediated exchange of α-synuclein oligomers between neighboring terminals could overwhelm intracellular trafficking by encumbering on the endosomal system. Much work still needs to be done to define the contribution of α-synuclein spreading to PD synaptopathy. However, the determination of what can be considered as a transmissible pathological form of α-synuclein as well as the mechanisms through which this entity can be transmitted from a diseased presynaptic terminal to a healthy postsynaptic ending can help us to understand much more on PD neurobiology and to identify novel effective therapeutic strategies to cure this disorder. Indeed, whether PD is primarily a disorder of the synapse, novel effective therapeutic approaches should both heal diseased synapses and block the cell-to-cell transmission of toxic α-synuclein species.

Table 1.

First author Title Year Journal
Evidences supporting α-synuclein spreading
Braak “Staging of Brain Pathology Related to Sporadic Parkinson's Disease” 2003 Neurobiol. Aging
Del Tredici “Sporadic Parkinson's Disease: Development and Distribution of Alpha-Synuclein Pathology” 2016 Neuropathol. Appl. Neurobiol.
Iljina “Kinetic Model of the Aggregation of Alpha-Synuclein Provides Insights into Prion-Like Spreading” 2016 Proc. Natl. Acad. Sci. USA
Oh “Mesenchymal Stem Cells Inhibit Transmission of α-Synuclein by Modulating Clathrin-Mediated Endocytosis in a Parkinsonian Model” 2016 Cell Rep.
Helwig “Brain Propagation of Transduced Alpha-Synuclein Involves Non-Fibrillar Protein Species and Is Enhanced in Alpha-Synuclein Null Mice” 2016 Brain
Bernis “Prion-Like Propagation of Human Brain-Derived Alpha-Synuclein in Transgenic Mice Expressing Human Wild-Type Alpha-Synuclein” 2015 Acta Neuropathol. Commun.
Illes-Toth “Distinct Higher-Order Alpha-Synuclein Oligomers Induce Intracellular Aggregation” 2015 Biochem. J.
Ulusoy “Neuron-to-Neuron α-Synuclein Propagation In Vivo Is Independent of Neuronal Injury” 2015 Acta Neuropathol. Commun.
Stuendl “Induction of α-Synuclein Aggregate Formation by CSF Exosomes from Patients with Parkinson's Disease and Dementia with Lewy Bodies” 2016 Brain
Prusiner “Evidence for α-Synuclein Prions Causing Multiple System Atrophy in Humans with Parkinsonism” 2015 Proc. Natl. Acad. Sci. USA
Masuda-Suzukake “Pathological Alpha-Synuclein Propagates through Neural Networks” 2014 Acta Neuropathol. Commun.
Sacino “Brain Injection of Alpha-Synuclein Induces Multiple Proteinopathies, Gliosis, and a Neuronal Injury Marker” 2014 J. Neurosci.
Kovacs “Intracellular Processing of Disease-Associated α-Synuclein in the Human Brain Suggests Prion-Like Cell-to-Cell Spread” 2014 Neurobiol. Dis.
Recasens “Lewy Body Extracts from Parkinson Disease Brains Trigger α-Synuclein Pathology and Neurodegeneration in Mice and Monkeys” 2014 Ann. Neurol.
Ulusoy “Caudo-Rostral Brain Spreading of α-Synuclein through Vagal Connections” 2013 EMBO Mol. Med.
Masuda-Suzukake “Prion-Like Spreading of Pathological α-Synuclein in Brain” 2013 Brain
Angot “Alpha-Synuclein Cell-to-Cell Transfer and Seeding in Grafted Dopaminergic Neurons In Vivo 2012 PLoS One
Luk “Intracerebral Inoculation of Pathological Alpha-Synuclein Initiates a Rapidly Progressive Neurodegenerative Alpha-Synucleinopathy in Mice” 2012 J. Exp. Med.
Luk “Pathological Alpha-Synuclein Transmission Initiates Parkinson-Like Neurodegeneration in Nontransgenic Mice” 2012 Science
Kordower “Transfer of Host-Derived Alpha Synuclein to Grafted Dopaminergic Neurons in Rat” 2011 Neurobiol. Dis.
Hansen “Alpha-Synuclein Propagates from Mouse Brain to Grafted Dopaminergic Neurons and Seeds Aggregation in Cultured Human Cells” 2011 J. Clin. Invest.
Danzer “Seeding Induced by Alpha-Synuclein Oligomers Provides Evidence for Spreading of Alpha-Synuclein Pathology” 2009 J. Neurochem.
Kordower “Transplanted Dopaminergic Neurons Develop PD Pathologic Changes: A Second Case Report” 2008 Mov. Disord.
Kordower “Lewy Body-Like Pathology in Long-Term Embryonic Nigral Transplants in Parkinson's Disease” 2008 Nat. Med.
Li “Lewy Bodies in Grafted Neurons in Subjects with Parkinson's Disease Suggest Host-to-Graft Disease Propagation” 2008 Nat. Med.

Evidences confuting α-synuclein spreading
Sumikura “Distribution of Alpha-Synuclein in the Spinal Cord and Dorsal Root Ganglia in an Autopsy Cohort of Elderly Persons” 2015 Acta Neuropathol. Commun.
Sacino “Amyloidogenic α-Synuclein Seeds Do Not Invariably Induce Rapid, Widespread Pathology in Mice” 2014 Acta Neuropathol.
Halliday “The Progression of Pathology in Parkinson's Disease” 2010 Ann. N. Y. Acad. Sci.
Jang “Non-Classical Exocytosis of Alpha-Synuclein Is Sensitive to Folding States and Promoted under Stress Conditions” 2010 J. Neurochem.
Hawkes “Parkinson's Disease and Aging: Same or Different Process?” 2008 Mov. Disord.
Kalaitzakis “Controversies over the Staging of Alpha-Synuclein Pathology in Parkinson's Disease” 2008 Acta Neuropathol.

Disclosure

PierFranco Spano is Professor Emeritus of Pharmacology, University of Brescia.

Competing Interests

The authors declare that there is no conflict of interests regarding the publication of this paper.

References

  • 1.Brose N., O'Connor V., Skehel P. Synaptopathy: dysfunction of synaptic function? Biochemical Society Transactions. 2010;38(2):443–444. doi: 10.1042/bst0380443. [DOI] [PubMed] [Google Scholar]
  • 2.Lepeta K., Lourenco M. V., Schweitzer B. C., et al. Synaptopathies: synaptic dysfunction in neurological disorders—a review from students to students. Journal of Neurochemistry. 2016;138(6):785–805. doi: 10.1111/jnc.13713. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Anichtchik O., Calo L., Spillantini M. G. Synaptic dysfunction in synucleinopathies. CNS and Neurological Disorders—Drug Targets. 2013;12(8):1094–1100. [PubMed] [Google Scholar]
  • 4.Zaltieri M., Grigoletto J., Longhena F., et al. α-synuclein and synapsin III cooperatively regulate synaptic function in dopamine neurons. Journal of Cell Science. 2015;128(13):2231–2243. doi: 10.1242/jcs.157867. [DOI] [PubMed] [Google Scholar]
  • 5.Calo L., Wegrzynowicz M., Santivañez-Perez J., Grazia Spillantini M. Synaptic failure and α-synuclein. Movement Disorders. 2016;31(2):169–177. doi: 10.1002/mds.26479. [DOI] [PubMed] [Google Scholar]
  • 6.Bellucci A., Mercuri N. B., Venneri A., et al. Review: Parkinson's disease: from synaptic loss to connectome dysfunction. Neuropathology and Applied Neurobiology. 2016;42(1):77–94. doi: 10.1111/nan.12297. [DOI] [PubMed] [Google Scholar]
  • 7.Bolam J. P., Pissadaki E. K. Living on the edge with too many mouths to feed: why dopamine neurons die. Movement Disorders. 2012;27(12):1478–1483. doi: 10.1002/mds.25135. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Aldecoa I., Navarro-Otano J., Stefanova N., et al. Alpha-synuclein immunoreactivity patterns in the enteric nervous system. Neuroscience Letters. 2015;602:145–149. doi: 10.1016/j.neulet.2015.07.005. [DOI] [PubMed] [Google Scholar]
  • 9.Recasens A., Dehay B., Bové J., et al. Lewy body extracts from Parkinson disease brains trigger α-synuclein pathology and neurodegeneration in mice and monkeys. Annals of Neurology. 2014;75(3):351–362. doi: 10.1002/ana.24066. [DOI] [PubMed] [Google Scholar]
  • 10.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. Nature Medicine. 2008;14(5):504–506. doi: 10.1038/nm1747. [DOI] [PubMed] [Google Scholar]
  • 11.Kordower J. H., Chu Y., Hauser R. A., Olanow C. W., Freeman T. B. Transplanted dopaminergic neurons develop PD pathologic changes: a second case report. Movement Disorders. 2008;23(16):2303–2306. doi: 10.1002/mds.22369. [DOI] [PubMed] [Google Scholar]
  • 12.Li W., Englund E., Widner H., et al. Extensive graft-derived dopaminergic innervation is maintained 24 years after transplantation in the degenerating parkinsonian brain. Proceedings of the National Academy of Sciences of the United States of America. 2016;113(23):6544–6549. doi: 10.1073/pnas.1605245113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Kovacs G. G., Breydo L., Green R., et al. Intracellular processing of disease-associated α-synuclein in the human brain suggests prion-like cell-to-cell spread. Neurobiology of Disease. 2014;69:76–92. doi: 10.1016/j.nbd.2014.05.020. [DOI] [PubMed] [Google Scholar]
  • 14.Braak H., Del Tredici K., Rüb U., De Vos R. A. I., Jansen Steur E. N. H., Braak E. Staging of brain pathology related to sporadic Parkinson's disease. Neurobiology of Aging. 2003;24(2):197–211. doi: 10.1016/S0197-4580(02)00065-9. [DOI] [PubMed] [Google Scholar]
  • 15.Del Tredici K., Braak H. Review: Sporadic Parkinson's disease: development and distribution of α-synuclein pathology. Neuropathology and Applied Neurobiology. 2016;42(1):33–50. doi: 10.1111/nan.12298. [DOI] [PubMed] [Google Scholar]
  • 16.George J. M. The synucleins. Genome Biology. 2002;3(1) doi: 10.1186/gb-2001-3-1-reviews3002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Ulusoy A., Di Monte D. A. α-Synuclein elevation in human neurodegenerative diseases: experimental, pathogenetic, and therapeutic implications. Molecular neurobiology. 2013;47(2):484–494. doi: 10.1007/s12035-012-8329-y. [DOI] [PubMed] [Google Scholar]
  • 18.Ulusoy A., Musgrove R. E., Rusconi R., et al. Neuron-to-neuron α-synuclein propagation in vivo is independent of neuronal injury. Acta Neuropathologica Communications. 2015;3(1, article 13) doi: 10.1186/s40478-015-0198-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Pan-Montojo F., Schwarz M., Winkler C., et al. Environmental toxins trigger PD-like progression via increased alpha-synuclein release from enteric neurons in mice. Scientific Reports. 2012;2, article 898 doi: 10.1038/srep00898. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Svensson E., Horváth-Puhó E., Thomsen R. W., et al. Vagotomy and subsequent risk of Parkinson's disease. Annals of Neurology. 2015;78(4):522–529. doi: 10.1002/ana.24448. [DOI] [PubMed] [Google Scholar]
  • 21.Brundin P., Ma J., Kordower J. H. How strong is the evidence that Parkinson's disease is a prion disorder? Current Opinion in Neurology. 2016;29(4):459–466. doi: 10.1097/wco.0000000000000349. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Bernis M. E., Babila J. T., Breid S., Wüsten K. A., Wüllner U., Tamgüney G. Prion-like propagation of human brain-derived alpha-synuclein in transgenic mice expressing human wild-type alpha-synuclein. Acta Neuropathologica Communications. 2015;3(1, article 75) doi: 10.1186/s40478-015-0254-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Breid S., Bernis M. E., Babila J. T., Garza M. C., Wille H., Tamgüney G. Neuroinvasion of α-synuclein prionoids after intraperitoneal and intraglossal inoculation. Journal of Virology. 2016;90(20):9182–9193. doi: 10.1128/jvi.01399-16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Prusiner S. B., Woerman A. L., Mordes D. A., et al. Evidence for α-synuclein prions causing multiple system atrophy in humans with parkinsonism. Proceedings of the National Academy of Sciences of the United States of America. 2015;112(38):E5308–E5317. doi: 10.1073/pnas.1514475112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Woerman A. L., Stöhr J., Aoyagi A., et al. Propagation of prions causing synucleinopathies in cultured cells. Proceedings of the National Academy of Sciences of the United States of America. 2015;112(35):E4949–E4958. doi: 10.1073/pnas.1513426112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Stuendl A., Kunadt M., Kruse N., et al. Induction of α-synuclein aggregate formation by CSF exosomes from patients with Parkinson’s disease and dementia with Lewy bodies. Brain. 2016;139(2):481–494. doi: 10.1093/brain/awv346. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Danzer K. M., Krebs S. K., Wolff M., Birk G., Hengerer B. Seeding induced by α-synuclein oligomers provides evidence for spreading of α-synuclein pathology. Journal of Neurochemistry. 2009;111(1):192–203. doi: 10.1111/j.1471-4159.2009.06324.x. [DOI] [PubMed] [Google Scholar]
  • 28.Angot E., Steiner J. A., Hansen C., Li J.-Y., Brundin P. Are synucleinopathies prion-like disorders? The Lancet Neurology. 2010;9(11):1128–1138. doi: 10.1016/s1474-4422(10)70213-1. [DOI] [PubMed] [Google Scholar]
  • 29.Olanow C. W., Brundin P. Parkinson's disease and alpha synuclein: is Parkinson's disease a prion-like disorder? Movement Disorders. 2013;28(1):31–40. doi: 10.1002/mds.25373. [DOI] [PubMed] [Google Scholar]
  • 30.Ugalde C. L., Finkelstein D. I., Lawson V. A., Hill A. F. Pathogenic mechanisms of prion protein, amyloid-β and α-synuclein misfolding: the prion concept and neurotoxicity of protein oligomers. Journal of Neurochemistry. 2016;139(2):162–180. doi: 10.1111/jnc.13772. [DOI] [PubMed] [Google Scholar]
  • 31.Iljina M., Garcia G. A., Horrocks M. H., et al. Kinetic model of the aggregation of alpha-synuclein provides insights into prion-like spreading. Proceedings of the National Academy of Sciences of the United States of America. 2016;113(9):E1206–E1215. doi: 10.1073/pnas.1524128113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Emmanouilidou E., Vekrellis K. Exocytosis and spreading of normal and aberrant α-synuclein. Brain Pathology. 2016;26(3):398–403. doi: 10.1111/bpa.12373. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Lee H.-J., Patel S., Lee S.-J. Intravesicular localization and exocytosis of α-synuclein and its aggregates. The Journal of Neuroscience. 2005;25(25):6016–6024. doi: 10.1523/jneurosci.0692-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Maroteaux L., Campanelli J. T., Scheller R. H. Synuclein: a neuron-specific protein localized to the nucleus and presynaptic nerve terminal. Journal of Neuroscience. 1988;8(8):2804–2815. doi: 10.1523/JNEUROSCI.08-08-02804.1988. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Wakabayashi K., Mori F., Tanji K., Orimo S., Takahashi H. Involvement of the peripheral nervous system in synucleinopathies, tauopathies and other neurodegenerative proteinopathies of the brain. Acta Neuropathologica. 2010;120(1):1–12. doi: 10.1007/s00401-010-0706-x. [DOI] [PubMed] [Google Scholar]
  • 36.Braak H., Del Tredici K. Invited Article: nervous system pathology in sporadic Parkinson disease. Neurology. 2008;70(20):1916–1925. doi: 10.1212/01.wnl.0000312279.49272.9f. [DOI] [PubMed] [Google Scholar]
  • 37.Böttner M., Fricke T., Müller M., et al. Alpha-synuclein is associated with the synaptic vesicle apparatus in the human and rat enteric nervous system. Brain Research. 2015;1614:51–59. doi: 10.1016/j.brainres.2015.04.015. [DOI] [PubMed] [Google Scholar]
  • 38.Totterdell S., Hanger D., Meredith G. E. The ultrastructural distribution of alpha-synuclein-like protein in normal mouse brain. Brain Research. 2004;1004(1-2):61–72. doi: 10.1016/j.brainres.2003.10.072. [DOI] [PubMed] [Google Scholar]
  • 39.Gordon S. L., Cousin M. A. The sybtraps: control of synaptobrevin traffic by synaptophysin, α-synuclein and AP-180. Traffic. 2014;15(3):245–254. doi: 10.1111/tra.12140. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Butler B., Saha K., Rana T., et al. Dopamine transporter activity is modulated by α-synuclein. The Journal of Biological Chemistry. 2015;290(49):29542–29554. doi: 10.1074/jbc.m115.691592. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Bellucci A., Collo G., Sarnico I., Battistin L., Missale C., Spano P. Alpha-synuclein aggregation and cell death triggered by energy deprivation and dopamine overload are counteracted by D2/D3 receptor activation. Journal of Neurochemistry. 2008;106(2):560–577. doi: 10.1111/j.1471-4159.2008.05406.x. [DOI] [PubMed] [Google Scholar]
  • 42.Bellucci A., Navarria L., Falarti E., et al. Redistribution of DAT/α-synuclein complexes visualized by ‘in situ’ proximity ligation assay in transgenic mice modelling early Parkinson's disease. PLoS ONE. 2011;6(12) doi: 10.1371/journal.pone.0027959.e27959 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Gitler A. D., Bevis B. J., Shorter J., et al. The Parkinson's disease protein α-synuclein disrupts cellular Rab homeostasis. Proceedings of the National Academy of Sciences of the United States of America. 2008;105(1):145–150. doi: 10.1073/pnas.0710685105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Chen Y., Lippincott-Schwartz J. Rab10 delivers GLUT4 storage vesicles to the plasma membrane. Communicative and Integrative Biology. 2013;6(3) doi: 10.4161/cib.23779.e23779 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Fusco G., De Simone A., Gopinath T., et al. Direct observation of the three regions in α-synuclein that determine its membrane-bound behaviour. Nature Communications. 2014;5, article 3827 doi: 10.1038/ncomms4827. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Fusco G., De Simone A., Arosio P., Vendruscolo M., Veglia G., Dobson C. M. Structural ensembles of membrane-bound α-synuclein reveal the molecular determinants of synaptic vesicle affinity. Scientific Reports. 2016;6 doi: 10.1038/srep27125.27125 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Shvadchak V. V., Yushchenko D. A., Pievo R., Jovin T. M. The mode of α-synuclein binding to membranes depends on lipid composition and lipid to protein ratio. FEBS Letters. 2011;585(22):3513–3519. doi: 10.1016/j.febslet.2011.10.006. [DOI] [PubMed] [Google Scholar]
  • 48.Kubo S.-I., Nemani V. M., Chalkley R. J., et al. A combinatorial code for the interaction of α-synuclein with membranes. The Journal of Biological Chemistry. 2005;280(36):31664–31672. doi: 10.1074/jbc.m504894200. [DOI] [PubMed] [Google Scholar]
  • 49.Burré J., Sharma M., Südhof T. C. Definition of a molecular pathway mediating α-synuclein neurotoxicity. Journal of Neuroscience. 2015;35(13):5221–5232. doi: 10.1523/JNEUROSCI.4650-14.2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Corbett G. T., Kordower J. H. The native form of α-Synuclein: monomer, tetramer, or a combination in equilibrium. Movement Disorders. 2015;30(14):1870–1870. doi: 10.1002/mds.26361. [DOI] [PubMed] [Google Scholar]
  • 51.Dettmer U., Newman A. J., Soldner F., et al. Parkinson-causing α-synuclein missense mutations shift native tetramers to monomers as a mechanism for disease initiation. Nature Communications. 2015;6:p. 7314. doi: 10.1038/ncomms8314. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Sahu K. K., Woodside M. T., Tuszynski J. A. α-synuclein dimer structures found from computational simulations. Biochimie. 2015;116:133–140. doi: 10.1016/j.biochi.2015.07.011. [DOI] [PubMed] [Google Scholar]
  • 53.Lorenzen N., Lemminger L., Pedersen J. N., Nielsen S. B., Otzen D. E. The N-terminus of α-synuclein is essential for both monomeric and oligomeric interactions with membranes. FEBS Letters. 2014;588(3):497–502. doi: 10.1016/j.febslet.2013.12.015. [DOI] [PubMed] [Google Scholar]
  • 54.Jiang Z., Hess S. K., Heinrich F., Lee J. C. Molecular details of α-synuclein membrane association revealed by neutrons and photons. Journal of Physical Chemistry B. 2015;119(14):4812–4823. doi: 10.1021/jp512499r. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Pirc K., Ulrih N. P. α-Synuclein interactions with phospholipid model membranes: key roles for electrostatic interactions and lipid-bilayer structure. Biochimica et Biophysica Acta—Biomembranes. 2015;1848(10):2002–2012. doi: 10.1016/j.bbamem.2015.06.021. [DOI] [PubMed] [Google Scholar]
  • 56.El-Agnaf O. M., Salem S. A., Paleologou K. E., et al. α-synuclein implicated in Parkinson's disease is present in extracellular biological fluids, including human plasma. The FASEB Journal. 2003;17(13):1945–1947. doi: 10.1096/fj.03-0098fje. [DOI] [PubMed] [Google Scholar]
  • 57.El-Agnaf O. M. A., Salem S. A., Paleologou K. E., et al. Detection of oligomeric forms of α-synuclein protein in human plasma as a potential biomarker for Parkinson's disease. The FASEB Journal. 2006;20(3):419–425. doi: 10.1096/fj.03-1449com. [DOI] [PubMed] [Google Scholar]
  • 58.Tokuda T., Salem S. A., Allsop D., et al. Decreased α-synuclein in cerebrospinal fluid of aged individuals and subjects with Parkinson's disease. Biochemical and Biophysical Research Communications. 2006;349(1):162–166. doi: 10.1016/j.bbrc.2006.08.024. [DOI] [PubMed] [Google Scholar]
  • 59.Devic I., Hwang H., Edgar J. S., et al. Salivary α-synuclein and DJ-1: potential biomarkers for Parkinson's disease. Brain. 2011;134:p. e178. doi: 10.1093/brain/awr015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Li J.-Y., Englund E., Holton J. L., et al. Lewy bodies in grafted neurons in subjects with Parkinson's disease suggest host-to-graft disease propagation. Nature Medicine. 2008;14(5):501–503. doi: 10.1038/nm1746. [DOI] [PubMed] [Google Scholar]
  • 61.Lee S. K., Sillitoe R. V., Silva C., Martina M., Sekerkova G. α-Synuclein expression in the mouse cerebellum is restricted to VGluT1 excitatory terminals and is enriched in unipolar brush cells. Cerebellum. 2015;14(5):516–527. doi: 10.1007/s12311-015-0673-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Jang A., Lee H.-J., Suk J.-E., Jung J.-W., Kim K.-P., Lee S.-J. Non-classical exocytosis of α-synuclein is sensitive to folding states and promoted under stress conditions. Journal of Neurochemistry. 2010;113(5):1263–1274. doi: 10.1111/j.1471-4159.2010.06695.x. [DOI] [PubMed] [Google Scholar]
  • 63.Emmanouilidou E., Melachroinou K., Roumeliotis T., et al. Cell-produced α-synuclein is secreted in a calcium-dependent manner by exosomes and impacts neuronal survival. The Journal of Neuroscience. 2010;30(20):6838–6851. doi: 10.1523/jneurosci.5699-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Alvarez-Erviti L., Seow Y., Schapira A. H., et al. Lysosomal dysfunction increases exosome-mediated alpha-synuclein release and transmission. Neurobiology of Disease. 2011;42(3):360–367. doi: 10.1016/j.nbd.2011.01.029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Stokholm M. G., Danielsen E. H., Hamilton-Dutoit S. J., Borghammer P. Pathological α-synuclein in gastrointestinal tissues from prodromal Parkinson disease patients. Annals of Neurology. 2016;79(6):940–949. doi: 10.1002/ana.24648. [DOI] [PubMed] [Google Scholar]
  • 66.Shannon K. M., Keshavarzian A., Mutlu E., et al. Alpha-synuclein in colonic submucosa in early untreated Parkinson's disease. Movement Disorders. 2012;27(6):709–715. doi: 10.1002/mds.23838. [DOI] [PubMed] [Google Scholar]
  • 67.Phillips R. J., Walter G. C., Wilder S. L., Baronowsky E. A., Powley T. L. Alpha-synuclein-immunopositive myenteric neurons and vagal preganglionic terminals: autonomic pathway implicated in Parkinson's disease? Neuroscience. 2008;153(3):733–750. doi: 10.1016/j.neuroscience.2008.02.074. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Donadio V., Incensi A., Leta V., et al. Skin nerve α-synuclein deposits: a biomarker for idiopathic Parkinson diseasee. Neurology. 2014;82(15):1362–1369. doi: 10.1212/wnl.0000000000000316. [DOI] [PubMed] [Google Scholar]
  • 69.Doppler K., Ebert S., Üçeyler N., et al. Cutaneous neuropathy in Parkinson's disease: a window into brain pathology. Acta Neuropathologica. 2014;128(1):99–109. doi: 10.1007/s00401-014-1284-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Gelpi E., Navarro-Otano J., Tolosa E., et al. Multiple organ involvement by alpha-synuclein pathology in lewy body disorders. Movement Disorders. 2014;29(8):1010–1018. doi: 10.1002/mds.25776. [DOI] [PubMed] [Google Scholar]
  • 71.Cersosimo M. G., Benarroch E. E. Pathological correlates of gastrointestinal dysfunction in Parkinson's disease. Neurobiology of Disease. 2012;46(3):559–564. doi: 10.1016/j.nbd.2011.10.014. [DOI] [PubMed] [Google Scholar]
  • 72.Mann D. M. A., Yates P. O. Pathological basis for neurotransmitter changes in Parkinson's disease. Neuropathology and Applied Neurobiology. 1983;9(1):3–19. doi: 10.1111/j.1365-2990.1983.tb00320.x. [DOI] [PubMed] [Google Scholar]
  • 73.Greene J. G. Causes and consequences of degeneration of the dorsal motor nucleus of the vagus nerve in Parkinson's disease. Antioxidants & Redox Signaling. 2014;21(4):649–667. doi: 10.1089/ars.2014.5859. [DOI] [PubMed] [Google Scholar]
  • 74.Sumikura H., Takao M., Hatsuta H., et al. Distribution of α-synuclein in the spinal cord and dorsal root ganglia in an autopsy cohort of elderly persons. Acta Neuropathologica Communications. 2015;3, article 57 doi: 10.1186/s40478-015-0236-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Hawkes C. H. Parkinson's disease and aging: same or different process? Movement Disorders. 2008;23(1):47–53. doi: 10.1002/mds.21766. [DOI] [PubMed] [Google Scholar]
  • 76.Buchman A. S., Shulman J. M., Nag S., et al. Nigral pathology and parkinsonian signs in elders without Parkinson disease. Annals of Neurology. 2012;71(2):258–266. doi: 10.1002/ana.22588. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Buchman A. S., Yu L., Wilson R. S., Schneider J. A., Bennett D. A. Association of brain pathology with the progression of frailty in older adults. Neurology. 2013;80(22):2055–2061. doi: 10.1212/WNL.0b013e318294b462. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Diederich N. J., Parent A. Parkinson's disease: acquired frailty of archaic neural networks? Journal of the Neurological Sciences. 2012;314(1-2):143–151. doi: 10.1016/j.jns.2011.10.003. [DOI] [PubMed] [Google Scholar]
  • 79.Kalaitzakis M. E., Graeber M. B., Gentleman S. M., Pearce R. K. B. The dorsal motor nucleus of the vagus is not an obligatory trigger site of Parkinson's disease: a critical analysis of α-synuclein staging. Neuropathology and Applied Neurobiology. 2008;34(3):284–295. doi: 10.1111/j.1365-2990.2007.00923.x. [DOI] [PubMed] [Google Scholar]
  • 80.Halliday G. M., McCann H. The progression of pathology in Parkinson's disease. Annals of the New York Academy of Sciences. 2010;1184:188–195. doi: 10.1111/j.1749-6632.2009.05118.x. [DOI] [PubMed] [Google Scholar]
  • 81.Wang X., Yu S., Li F., Feng T. Detection of α-synuclein oligomers in red blood cells as a potential biomarker of Parkinson's disease. Neuroscience Letters. 2015;599:115–119. doi: 10.1016/j.neulet.2015.05.030. [DOI] [PubMed] [Google Scholar]
  • 82.Tokuda T., Qureshi M. M., Ardah M. T., et al. Detection of elevated levels of α-synuclein oligomers in CSF from patients with Parkinson disease. Neurology. 2010;75(20):1766–1772. doi: 10.1212/wnl.0b013e3181fd613b. [DOI] [PubMed] [Google Scholar]
  • 83.Kim S., Jeon B. S., Heo C., et al. α-Synuclein induces apoptosis by altered expression in human peripheral lymphocyte in Parkinson's disease. FASEB Journal. 2004;18(13):1615–1617. doi: 10.1096/fj.04-1917fje. [DOI] [PubMed] [Google Scholar]
  • 84.Shi M., Liu C., Cook T. J., et al. Plasma exosomal α-synuclein is likely CNS-derived and increased in Parkinson's disease. Acta Neuropathologica. 2014;128(5):639–650. doi: 10.1007/s00401-014-1314-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Joshi G. K., Deitz-McElyea S., Liyanage T., et al. Label-free nanoplasmonic-based short noncoding RNA sensing at attomolar concentrations allows for quantitative and highly specific assay of microRNA-10b in biological fluids and circulating exosomes. ACS Nano. 2015;9(11):11075–11089. doi: 10.1021/acsnano.5b04527. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Stuendl A., Kunadt M., Kruse N., et al. Induction of α-synuclein aggregate formation by CSF exosomes from patients with Parkinson's disease and dementia with Lewy bodies. Brain. 2016;139, part 2:481–494. doi: 10.1093/brain/awv346. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Salahuddin P., Fatima M. T., Abdelhameed A. S., Nusrat S., Khan R. H. ChemInform abstract: structure of amyloid oligomers and their mechanisms of toxicities: targeting amyloid oligomers using novel therapeutic approaches. ChemInform. 2016;47(22) doi: 10.1002/chin.201622261. [DOI] [PubMed] [Google Scholar]
  • 88.Villar-Piqué A., da Fonseca T. L., Outeiro T. F. Structure, function and toxicity of alpha-synuclein: the Bermuda triangle in synucleinopathies. Journal of Neurochemistry. 2016;139(1):240–255. doi: 10.1111/jnc.13249. [DOI] [PubMed] [Google Scholar]
  • 89.Peelaerts W., Bousset L., Van der Perren A., et al. α-Synuclein strains cause distinct synucleinopathies after local and systemic administration. Nature. 2015;522(7556):340–344. doi: 10.1038/nature14547. [DOI] [PubMed] [Google Scholar]
  • 90.Park J.-Y., Paik S. R., Jou I., Park S. M. Microglial phagocytosis is enhanced by monomeric α-synuclein, not aggregated α-synuclein: implications for Parkinson's disease. Glia. 2008;56(11):1215–1223. doi: 10.1002/glia.20691. [DOI] [PubMed] [Google Scholar]
  • 91.Kim K. S., Kim J. S., Park J.-Y., et al. DJ-1 associates with lipid rafts by palmitoylation and regulates lipid rafts-dependent endocytosis in astrocytes. Human Molecular Genetics. 2013;22(23):4805–4817. doi: 10.1093/hmg/ddt332. [DOI] [PubMed] [Google Scholar]
  • 92.Hoffmann A., Ettle B., Bruno A., et al. Alpha-synuclein activates BV2 microglia dependent on its aggregation state. Biochemical and Biophysical Research Communications. 2016;479(4):881–886. doi: 10.1016/j.bbrc.2016.09.109. [DOI] [PubMed] [Google Scholar]
  • 93.Zhang W., Wang T., Pei Z., et al. Aggregated α-synuclein activates microglia: a process leading to disease progression in Parkinson's disease. The FASEB Journal. 2005;19(6):533–542. doi: 10.1096/fj.04-2751com. [DOI] [PubMed] [Google Scholar]
  • 94.Lee J. H., Ying J., Bax A. Nuclear magnetic resonance observation of α-synuclein membrane interaction by monitoring the acetylation reactivity of its lysine side chains. Biochemistry. 2016;55(35):4949–4959. doi: 10.1021/acs.biochem.6b00637. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Salveson P. J., Spencer R. K., Nowick J. S. X-ray crystallographic structure of oligomers formed by a toxic β-hairpin derived from α-synuclein: trimers and higher-order oligomers. Journal of the American Chemical Society. 2016;138(13):4458–4467. doi: 10.1021/jacs.5b13261. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Chen S. W., Drakulic S., Deas E., et al. Structural characterization of toxic oligomers that are kinetically trapped during α-synuclein fibril formation. Proceedings of the National Academy of Sciences of the United States of America. 2015;112(16):E1994–E2003. doi: 10.1073/pnas.1421204112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Roostaee A., Beaudoin S., Staskevicius A., Roucou X. Aggregation and neurotoxicity of recombinant α-synuclein aggregates initiated by dimerization. Molecular Neurodegeneration. 2013;8, article 5 doi: 10.1186/1750-1326-8-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Tsigelny I. F., Sharikov Y., Wrasidlo W., et al. Role of α-synuclein penetration into the membrane in the mechanisms of oligomer pore formation. FEBS Journal. 2012;279(6):1000–1013. doi: 10.1111/j.1742-4658.2012.08489.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Winner B., Jappelli R., Maji S. K., et al. In vivo demonstration that α-synuclein oligomers are toxic. Proceedings of the National Academy of Sciences of the United States of America. 2011;108(10):4194–4199. doi: 10.1073/pnas.1100976108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Celej M. S., Sarroukh R., Goormaghtigh E., Fidelio G. D., Ruysschaert J.-M., Raussens V. Toxic prefibrillar α-synuclein amyloid oligomers adopt a distinctive antiparallel β-sheet structure. Biochemical Journal. 2012;443(3):719–726. doi: 10.1042/BJ20111924. [DOI] [PubMed] [Google Scholar]
  • 101.Plotegher N., Gratton E., Bubacco L. Number and Brightness analysis of alpha-synuclein oligomerization and the associated mitochondrial morphology alterations in live cells. Biochimica et Biophysica Acta (BBA)—General Subjects. 2014;1840(6):2014–2024. doi: 10.1016/j.bbagen.2014.02.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.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. Neurobiology of Disease. 2014;70:204–213. doi: 10.1016/j.nbd.2014.06.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Luth E. S., Stavrovskaya I. G., Bartels T., Kristal B. S., Selkoe D. J. Soluble, prefibrillar α-synuclein oligomers promote complex I-dependent, Ca2+-induced mitochondrial dysfunction. Journal of Biological Chemistry. 2014;289(31):21490–21507. doi: 10.1074/jbc.M113.545749. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Stefanovic A. N. D., Stöckl M. T., Claessens M. M. A. E., Subramaniam V. α-Synuclein oligomers distinctively permeabilize complex model membranes. The FEBS Journal. 2014;281(12):2838–2850. doi: 10.1111/febs.12824. [DOI] [PubMed] [Google Scholar]
  • 105.Ryan B. J., Hoek S., Fon E. A., Wade-Martins R. Mitochondrial dysfunction and mitophagy in Parkinson's: from familial to sporadic disease. Trends in Biochemical Sciences. 2015;40(4):200–210. doi: 10.1016/j.tibs.2015.02.003. [DOI] [PubMed] [Google Scholar]
  • 106.Mullin S., Schapira A. α-Synuclein and mitochondrial dysfunction in Parkinson's disease. Molecular neurobiology. 2013;47(2):587–597. doi: 10.1007/s12035-013-8394-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Zaltieri M., Longhena F., Pizzi M., Missale C., Spano P., Bellucci A. Mitochondrial dysfunction and α-synuclein synaptic pathology in Parkinson's disease: who's on first? Parkinson's Disease. 2015;2015:10. doi: 10.1155/2015/108029.108029 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Colla E., Jensen P. H., Pletnikova O., Troncoso J. C., Glabe C., Lee M. K. Accumulation of toxic α-synuclein oligomer within endoplasmic reticulum occurs in α-synucleinopathy in vivo. Journal of Neuroscience. 2012;32(10):3301–3305. doi: 10.1523/jneurosci.5368-11.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Bellucci A., Navarria L., Zaltieri M., et al. Induction of the unfolded protein response by α-synuclein in experimental models of Parkinson's disease. Journal of Neurochemistry. 2011;116(4):588–605. doi: 10.1111/j.1471-4159.2010.07143.x. [DOI] [PubMed] [Google Scholar]
  • 110.Deas E., Cremades N., Angelova P. R., et al. Alpha-synuclein oligomers interact with metal ions to induce oxidative stress and neuronal death in Parkinson's disease. Antioxidants & Redox Signaling. 2016;24(7):376–391. doi: 10.1089/ars.2015.6343. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Rockenstein E., Nuber S., Overk C. R., et al. Accumulation of oligomer-prone α-synuclein exacerbates synaptic and neuronal degeneration in vivo. Brain. 2014;137(5):1496–1513. doi: 10.1093/brain/awu057. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Kaufmann T. J., Harrison P. M., Richardson M. J., Pinheiro T. J., Wall M. J. Intracellular soluble α-synuclein oligomers reduce pyramidal cell excitability. The Journal of Physiology. 2016;594(10):2751–2772. doi: 10.1113/JP271968. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Diógenes M. J., Dias R. B., Rombo D. M., et al. Extracellular alpha-synuclein oligomers modulate synaptic transmission and impair LTP via NMDA-receptor activation. Journal of Neuroscience. 2012;32(34):11750–11762. doi: 10.1523/JNEUROSCI.0234-12.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Emanuele M., Esposito A., Camerini S., et al. Exogenous alpha-synuclein alters pre- and post-synaptic activity by fragmenting lipid rafts. EBioMedicine. 2016;7:191–204. doi: 10.1016/j.ebiom.2016.03.038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Roberts H. L., Brown D. R. Seeking a mechanism for the toxicity of oligomeric α-synuclein. Biomolecules. 2015;5(2):282–305. doi: 10.3390/biom5020282. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Fearnley J. M., Lees A. J. Ageing and Parkinson's disease: substantia nigra regional selectivity. Brain. 1991;114(5):2283–2301. doi: 10.1093/brain/114.5.2283. [DOI] [PubMed] [Google Scholar]
  • 117.Halliday G. M., McRitchie D. A., Cartwright H., Pamphlett R., Hely M. A., Morris J. G. L. Midbrain neuropathology in idiopathic Parkinson's disease and diffuse Lewy body disease. Journal of Clinical Neuroscience. 1996;3(1):52–60. doi: 10.1016/S0967-5868(96)90083-1. [DOI] [PubMed] [Google Scholar]
  • 118.Venda L. L., Cragg S. J., Buchman V. L., Wade-Martins R. α-Synuclein and dopamine at the crossroads of Parkinson's disease. Trends in Neurosciences. 2010;33(12):559–568. doi: 10.1016/j.tins.2010.09.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Conway K. A., Rochet J.-C., Bieganski R. M., Lansbury P. T., Jr. Kinetic stabilization of the α-synuclein protofibril by a dopamine-α-synuclein adduct. Science. 2001;294(5545):1346–1349. doi: 10.1126/science.1063522. [DOI] [PubMed] [Google Scholar]
  • 120.Rekas A., Knott R. B., Sokolova A., et al. The structure of dopamine induced α-synuclein oligomers. European Biophysics Journal. 2010;39(10):1407–1419. doi: 10.1007/s00249-010-0595-x. [DOI] [PubMed] [Google Scholar]
  • 121.Follmer C., Coelho-Cerqueira E., Yatabe-Franco D. Y., et al. Oligomerization and membrane-binding properties of covalent adducts formed by the interaction of α-synuclein with the toxic dopamine metabolite 3,4-dihydroxyphenylacetaldehyde (DOPAL) Journal of Biological Chemistry. 2015;290(46):27660–27679. doi: 10.1074/jbc.M115.686584. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Curtain C. C., Kirby N. M., Mertens H. D. T., et al. Alpha-synuclein oligomers and fibrils originate in two distinct conformer pools: A Small Angle X-Ray Scattering And Ensemble Optimisation Modelling Study. Molecular BioSystems. 2015;11(1):190–196. doi: 10.1039/c4mb00356j. [DOI] [PubMed] [Google Scholar]
  • 123.Lemkau L. R., Comellas G., Lee S. W., et al. Site-specific perturbations of alpha-synuclein fibril structure by the Parkinson's disease associated mutations A53T and E46K. PLoS ONE. 2013;8(3) doi: 10.1371/journal.pone.0049750.e49750 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Harada R., Kobayashi N., Kim J., et al. The effect of amino acid substitution in the imperfect repeat sequences of α-synuclein on fibrillation. Biochimica et Biophysica Acta. 2009;1792(10):998–1003. doi: 10.1016/j.bbadis.2009.06.010. [DOI] [PubMed] [Google Scholar]
  • 125.Kelly J. W. The alternative conformations of amyloidogenic proteins and their multi-step assembly pathways. Current Opinion in Structural Biology. 1998;8(1):101–106. doi: 10.1016/S0959-440X(98)80016-X. [DOI] [PubMed] [Google Scholar]
  • 126.Ono K., Yamada M. Low-n oligomers as therapeutic targets of Alzheimer's disease. Journal of Neurochemistry. 2011;117(1):19–28. doi: 10.1111/j.1471-4159.2011.07187.x. [DOI] [PubMed] [Google Scholar]
  • 127.Serpell L. C., Berriman J., Jakes R., Goedert M., Crowther R. A. Fiber diffraction of synthetic α-synuclein filaments shows amyloid-like cross-β conformation. Proceedings of the National Academy of Sciences of the United States of America. 2000;97(9):4897–4902. doi: 10.1073/pnas.97.9.4897. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Shvadchak V. V., Subramaniam V. A four-amino acid linker between repeats in the α-synuclein sequence is important for fibril formation. Biochemistry. 2014;53(2):279–281. doi: 10.1021/bi401427t. [DOI] [PubMed] [Google Scholar]
  • 129.Izawa Y., Tateno H., Kameda H., et al. Role of C-terminal negative charges and tyrosine residues in fibril formation of α-synuclein. Brain and Behavior. 2012;2(5):595–605. doi: 10.1002/brb3.86. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Levitan K., Chereau D., Cohen S. I. A., et al. Conserved C-terminal charge exerts a profound influence on the aggregation rate of α-synuclein. Journal of Molecular Biology. 2011;411(2):329–333. doi: 10.1016/j.jmb.2011.05.046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Jo E., McLaurin J., Yip C. M., St George-Hyslop P., Fraser P. E. α-Synuclein membrane interactions and lipid specificity. Journal of Biological Chemistry. 2000;275(44):34328–34334. doi: 10.1074/jbc.m004345200. [DOI] [PubMed] [Google Scholar]
  • 132.Tsigelny I. F., Sharikov Y., Miller M. A., Masliah E. Mechanism of alpha-synuclein oligomerization and membrane interaction: theoretical approach to unstructured proteins studies. Nanomedicine: Nanotechnology, Biology, and Medicine. 2008;4(4):350–357. doi: 10.1016/j.nano.2008.05.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Danzer K. M., Haasen D., Karow A. R., et al. Different species of α-synuclein oligomers induce calcium influx and seeding. The Journal of Neuroscience. 2007;27(34):9220–9232. doi: 10.1523/jneurosci.2617-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Caughey B., Lansbury P. T., Jr. Protofibrils, pores, fibrils, and neurodegeneration: separating the responsible protein aggregates from the innocent bystanders. Annual Review of Neuroscience. 2003;26:267–298. doi: 10.1146/annurev.neuro.26.010302.081142. [DOI] [PubMed] [Google Scholar]
  • 135.Lam H. T., Graber M. C., Gentry K. A., Bieschke J. Stabilization of α-synuclein fibril clusters prevents fragmentation and reduces seeding activity and toxicity. Biochemistry. 2016;55(4):675–685. doi: 10.1021/acs.biochem.5b01168. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Narayanan V., Scarlata S. Membrane binding and self-association of α-synucleins. Biochemistry. 2001;40(33):9927–9934. doi: 10.1021/bi002952n. [DOI] [PubMed] [Google Scholar]
  • 137.Lee H.-J., Choi C., Lee S.-J. Membrane-bound α-synuclein has a high aggregation propensity and the ability to seed the aggregation of the cytosolic form. The Journal of Biological Chemistry. 2002;277(1):671–678. doi: 10.1074/jbc.m107045200. [DOI] [PubMed] [Google Scholar]
  • 138.Burré J. The synaptic function of α-synuclein. Journal of Parkinson's Disease. 2015;5(4):699–713. doi: 10.3233/JPD-150642. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Zhu M., Fink A. L. Lipid binding inhibits α-synuclein fibril formation. Journal of Biological Chemistry. 2003;278(19):16873–16877. doi: 10.1074/jbc.M210136200. [DOI] [PubMed] [Google Scholar]
  • 140.Hansen C., Angot E., Bergström A.-L., et al. α-Synuclein propagates from mouse brain to grafted dopaminergic neurons and seeds aggregation in cultured human cells. Journal of Clinical Investigation. 2011;121(2):715–725. doi: 10.1172/JCI43366. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Kordower J. H., Dodiya H. B., Kordower A. M., et al. Transfer of host-derived alpha synuclein to grafted dopaminergic neurons in rat. Neurobiology of Disease. 2011;43(3):552–557. doi: 10.1016/j.nbd.2011.05.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Angot E., Steiner J. A., Tomé C. M., et al. α-synuclein cell-to-cell transfer and seeding in grafted dopaminergic neurons in vivo. PLoS ONE. 2012;7(6) doi: 10.1371/journal.pone.0039465.e39465 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Dixon C., Mathias N., Zweig R. M., Davis D. A., Gross D. S. α-synuclein targets the plasma membrane via the secretory pathway and induces toxicity in yeast. Genetics. 2005;170(1):47–59. doi: 10.1534/genetics.104.035493. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Rubinsztein D. C. The roles of intracellular protein-degradation pathways in neurodegeneration. Nature. 2006;443(7113):780–786. doi: 10.1038/nature05291. [DOI] [PubMed] [Google Scholar]
  • 145.Pan T., Kondo S., Le W., Jankovic J. The role of autophagy-lysosome pathway in neurodegeneration associated with Parkinson's disease. Brain. 2008;131(8):1969–1978. doi: 10.1093/brain/awm318. [DOI] [PubMed] [Google Scholar]
  • 146.Tofaris G. K., Layfield R., Spillantini M. G. α-Synuclein metabolism and aggregation is linked to ubiquitin-independent degradation by the proteasome. FEBS Letters. 2001;509(1):22–26. doi: 10.1016/S0014-5793(01)03115-5. [DOI] [PubMed] [Google Scholar]
  • 147.Larsen K. E., Sulzer D. Autophagy in neurons: a review. Histology and Histopathology. 2002;17(3):897–908. doi: 10.14670/HH-17.897. [DOI] [PubMed] [Google Scholar]
  • 148.McNaught K. S. P., Belizaire R., Isacson O., Jenner P., Olanow C. W. Altered proteasomal function in sporadic Parkinson's disease. Experimental Neurology. 2003;179(1):38–46. doi: 10.1006/exnr.2002.8050. [DOI] [PubMed] [Google Scholar]
  • 149.Olanow C. W., McNaught K. S. P. Ubiquitin-proteasome system and Parkinson's disease. Movement Disorders. 2006;21(11):1806–1823. doi: 10.1002/mds.21013. [DOI] [PubMed] [Google Scholar]
  • 150.Tofaris G. K., Razzaq A., Ghetti B., Lilley K. S., Spillantini M. G. Ubiquitination of α-synuclein in lewy bodies is a pathological event not associated with impairment of proteasome function. Journal of Biological Chemistry. 2003;278(45):44405–44411. doi: 10.1074/jbc.m308041200. [DOI] [PubMed] [Google Scholar]
  • 151.Cuervo A. M. Autophagy: many paths to the same end. Molecular and Cellular Biochemistry. 2004;263(1):55–72. doi: 10.1023/b:mcbi.0000041848.57020.57. [DOI] [PubMed] [Google Scholar]
  • 152.Meredith G. E., Totterdell S., Petroske E., Santa Cruz K., Callison R. C., Jr., Lau Y.-S. Lysosomal malfunction accompanies alpha-synuclein aggregation in a progressive mouse model of Parkinson's disease. Brain Research. 2002;956(1):156–165. doi: 10.1016/s0006-8993(02)03514-x. [DOI] [PubMed] [Google Scholar]
  • 153.Keller J. N., Dimayuga E., Chen Q., Thorpe J., Gee J., Ding Q. Autophagy, proteasomes, lipofuscin, and oxidative stress in the aging brain. International Journal of Biochemistry and Cell Biology. 2004;36(12):2376–2391. doi: 10.1016/j.biocel.2004.05.003. [DOI] [PubMed] [Google Scholar]
  • 154.Martinez-Vicente M., Sovak G., Cuervo A. M. Protein degradation and aging. Experimental Gerontology. 2005;40(8-9):622–633. doi: 10.1016/j.exger.2005.07.005. [DOI] [PubMed] [Google Scholar]
  • 155.Kiffin R., Kaushik S., Zeng M., et al. Altered dynamics of the lysosomal receptor for chaperone-mediated autophagy with age. Journal of Cell Science. 2007;120(5):782–791. doi: 10.1242/jcs.001073. [DOI] [PubMed] [Google Scholar]
  • 156.Reyes J. F., Olsson T. T., Lamberts J. T., Devine M. J., Kunath T., Brundin P. A cell culture model for monitoring α-synuclein cell-to-cell transfer. Neurobiology of Disease. 2015;77:266–275. doi: 10.1016/j.nbd.2014.07.003. [DOI] [PubMed] [Google Scholar]
  • 157.Alvarez-Erviti L., Seow Y., Yin H., Betts C., Lakhal S., Wood M. J. A. Delivery of siRNA to the mouse brain by systemic injection of targeted exosomes. Nature Biotechnology. 2011;29(4):341–345. doi: 10.1038/nbt.1807. [DOI] [PubMed] [Google Scholar]
  • 158.Liu J., Zhang J.-P., Shi M., et al. Rab11a and HSP90 regulate recycling of extracellular α-synuclein. Journal of Neuroscience. 2009;29(5):1480–1485. doi: 10.1523/JNEUROSCI.6202-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Verweij F. J., Middeldorp J. M., Pegtel D. M. Intracellular signaling controlled by the endosomal-exosomal pathway. Communicative & Integrative Biology. 2014;5(1):88–93. doi: 10.4161/cib.18452. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Kunadt M., Eckermann K., Stuendl A., et al. Extracellular vesicle sorting of α-Synuclein is regulated by sumoylation. Acta Neuropathologica. 2015;129(5):695–713. doi: 10.1007/s00401-015-1408-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Lo Cicero A., Stahl P. D., Raposo G. Extracellular vesicles shuffling intercellular messages: for good or for bad. Current Opinion in Cell Biology. 2015;35:69–77. doi: 10.1016/j.ceb.2015.04.013. [DOI] [PubMed] [Google Scholar]
  • 162.Théry C., Zitvogel L., Amigorena S. Exosomes: composition, biogenesis and function. Nature Reviews Immunology. 2002;2(8):569–579. doi: 10.1038/nri855. [DOI] [PubMed] [Google Scholar]
  • 163.Hasegawa T., Konno M., Baba T., et al. The AAA-ATPase VPS4 regulates extracellular secretion and lysosomal targeting of α-synuclein. PLoS ONE. 2011;6(12) doi: 10.1371/journal.pone.0029460.e29460 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Luk K. C., Kehm V., Carroll J., et al. Pathological α-synuclein transmission initiates Parkinson-like neurodegeneration in nontransgenic mice. Science. 2012;338(6109):949–953. doi: 10.1126/science.1227157. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165.Marblestone J. G., Edavettal S. C., Lim Y., Lim P., Zuo X., Butt T. R. Comparison of SUMO fusion technology with traditional gene fusion systems: enhanced expression and solubility with SUMO. Protein Science. 2006;15(1):182–189. doi: 10.1110/ps.051812706. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Panavas T., Sanders C., Butt T. R. SUMO fusion technology for enhanced protein production in prokaryotic and eukaryotic expression systems. Methods in Molecular Biology. 2009;497:303–317. doi: 10.1007/978-1-59745-566-4_20. [DOI] [PubMed] [Google Scholar]
  • 167.Chung K. K. K., Zhang Y., Lim K. L., et al. Parkin ubiquitinates the α-synuclein−interacting protein, synphilin-1: implications for Lewy-body formation in Parkinson disease. Nature Medicine. 2001;7(10):1144–1150. doi: 10.1038/nm1001-1144. [DOI] [PubMed] [Google Scholar]
  • 168.Sharma N., McLean P. J., Kawamata H., Irizarry M. C., Hyman B. T. Alpha-synuclein has an altered conformation and shows a tight intermolecular interaction with ubiquitin in Lewy bodies. Acta Neuropathologica. 2001;102(4):329–334. doi: 10.1007/s004010100369. [DOI] [PubMed] [Google Scholar]
  • 169.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 α-synuclein. International Journal of Biochemistry and Cell Biology. 2008;40(9):1835–1849. doi: 10.1016/j.biocel.2008.01.017. [DOI] [PubMed] [Google Scholar]
  • 170.Rott R., Szargel R., Haskin J., et al. Monoubiquitylation of α-synuclein by seven in absentia homolog (SIAH) promotes its aggregation in dopaminergic cells. Journal of Biological Chemistry. 2008;283(6):3316–3328. doi: 10.1074/jbc.m704809200. [DOI] [PubMed] [Google Scholar]
  • 171.Schulz-Schaeffer W. J. The synaptic pathology of α-synuclein aggregation in dementia with Lewy bodies, Parkinson's disease and Parkinson's disease dementia. Acta Neuropathologica. 2010;120(2):131–143. doi: 10.1007/s00401-010-0711-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Luk K. C., Kehm V. M., Zhang B., O'Brien P., Trojanowski J. Q., Lee V. M. Y. Intracerebral inoculation of pathological α-synuclein initiates a rapidly progressive neurodegenerative α-synucleinopathy in mice. Journal of Experimental Medicine. 2012;209(5):975–988. doi: 10.1084/jem.20112457. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Masuda-Suzukake M., Nonaka T., Hosokawa M., et al. Prion-like spreading of pathological α-synuclein in brain. Brain. 2013;136, part 4:1128–1138. doi: 10.1093/brain/awt037. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174.Masuda-Suzukake M., Nonaka T., Hosokawa M., et al. Pathological alpha-synuclein propagates through neural networks. Acta Neuropathologica Communications. 2014;2, article no. 88 doi: 10.1186/preaccept-1296467154135944. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175.Sacino A. N., Brooks M., McKinney A. B., et al. Brain injection of α-synuclein induces multiple proteinopathies, gliosis, and a neuronal injury marker. The Journal of Neuroscience. 2014;34(37):12368–12378. doi: 10.1523/jneurosci.2102-14.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176.Tarutani A., Suzuki G., Shimozawa A., et al. The effect of fragmented pathogenic α-synuclein seeds on prion-like propagation. Journal of Biological Chemistry. 2016;291(36):18675–18688. doi: 10.1074/jbc.m116.734707. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Helwig M., Klinkenberg M., Rusconi R., et al. Brain propagation of transduced α-synuclein involves non-fibrillar protein species and is enhanced in α-synuclein null mice. Brain. 2016;139, part 3:856–870. doi: 10.1093/brain/awv376. [DOI] [PubMed] [Google Scholar]
  • 178.Danzer K. M., Kranich L. R., Ruf W. P., et al. Exosomal cell-to-cell transmission of alpha synuclein oligomers. Molecular Neurodegeneration. 2012;7(1, article 42) doi: 10.1186/1750-1326-7-42. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Illes-Toth E., Ramos M. R., Cappai R., Dalton C., Smith D. P. Distinct higher-order α -synuclein oligomers induce intracellular aggregation. Biochemical Journal. 2015;468(3):485–493. doi: 10.1042/BJ20150159. [DOI] [PubMed] [Google Scholar]
  • 180.Chadchankar H., Ihalainen J., Tanila H., Yavich L. Decreased reuptake of dopamine in the dorsal striatum in the absence of alpha-synuclein. Brain Research. 2011;1382:37–44. doi: 10.1016/j.brainres.2011.01.064. [DOI] [PubMed] [Google Scholar]
  • 181.Liu S., Fa M., Ninan I., Trinchese F., Dauer W., Arancio O. α-Synuclein involvement in hippocampal synaptic plasticity: role of NO, cGMP, cGK and CaMKII. European Journal of Neuroscience. 2007;25(12):3583–3596. doi: 10.1111/j.1460-9568.2007.05569.x. [DOI] [PubMed] [Google Scholar]
  • 182.Bellucci A., Navarria L., Zaltieri M., Missale C., Spano P. Alpha-synuclein synaptic pathology and its implications in the development of novel therapeutic approaches to cure Parkinson's disease. Brain Research. 2012;1432:95–113. doi: 10.1016/j.brainres.2011.11.031. [DOI] [PubMed] [Google Scholar]
  • 183.Dikiy I., Eliezer D. Folding and misfolding of alpha-synuclein on membranes. Biochimica et Biophysica Acta—Biomembranes. 2012;1818(4):1013–1018. doi: 10.1016/j.bbamem.2011.09.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184.Chaudhary H., Stefanovic A. N. D., Subramaniam V., Claessens M. M. A. E. Membrane interactions and fibrillization of α-synuclein play an essential role in membrane disruption. FEBS Letters. 2014;588(23):4457–4463. doi: 10.1016/j.febslet.2014.10.016. [DOI] [PubMed] [Google Scholar]
  • 185.Bodner C. R., Maltsev A. S., Dobson C. M., Bax A. Differential phospholipid binding of α-synuclein variants implicated in Parkinson's disease revealed by solution NMR spectroscopy. Biochemistry. 2010;49(5):862–871. doi: 10.1021/bi901723p. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Stefanovic A. N. D., Claessens M. M. A. E., Blum C., Subramaniam V. Alpha-synuclein amyloid oligomers act as multivalent nanoparticles to cause hemifusion in negatively charged vesicles. Small. 2015;11(19):2257–2262. doi: 10.1002/smll.201402674. [DOI] [PubMed] [Google Scholar]
  • 187.Gonzalez-Horta A. The interaction of alpha-synuclein with membranes and its implication in parkinsońs disease: a literature review. Natural Product Communications. 2015;10(10):1775–1778. [PubMed] [Google Scholar]
  • 188.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 α-synuclein. Journal of Biological Chemistry. 2001;276(29):27441–27448. doi: 10.1074/jbc.m101318200. [DOI] [PubMed] [Google Scholar]
  • 189.Holmes B. B., DeVos S. L., Kfoury N., et al. Heparan sulfate proteoglycans mediate internalization and propagation of specific proteopathic seeds. Proceedings of the National Academy of Sciences of the United States of America. 2013;110(33):E3138–E3147. doi: 10.1073/pnas.1301440110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190.Madeira A., Yang J., Zhang X., et al. Caveolin-1 interacts with alpha-synuclein and mediates toxic actions of cellular alpha-synuclein overexpression. Neurochemistry International. 2011;59(2):280–289. doi: 10.1016/j.neuint.2011.05.017. [DOI] [PubMed] [Google Scholar]
  • 191.Oh S. H., Kim H. N., Park H. J., Shin J. Y., Kim D. Y., Lee P. H. The cleavage effect of mesenchymal stem cell and its derived matrix metalloproteinase-2 on extracellular α-synuclein aggregates in Parkinsonian models. Stem Cells Translational Medicine. 2016 doi: 10.5966/sctm.2016-0111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.Mao X., Ou M. T., Karuppagounder S. S., et al. Pathological α-synuclein transmission initiated by binding lymphocyte-activation gene 3. Science. 2016;353(6307) doi: 10.1126/science.aah3374.aah3374 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 193.Li K., Reichmann H., Ziemssen T. Recognition and treatment of autonomic disturbances in Parkinsons disease. Expert Review of Neurotherapeutics. 2015;15(10):1189–1203. doi: 10.1586/14737175.2015.1095093. [DOI] [PubMed] [Google Scholar]

Articles from Neural Plasticity are provided here courtesy of Wiley

RESOURCES