Next Article in Journal
Investigation of the Presence of DNA in Human Blood Plasma Small Extracellular Vesicles
Next Article in Special Issue
Revisiting Alpha-Synuclein Pathways to Inflammation
Previous Article in Journal
Retinal Organoids from an AIPL1 CRISPR/Cas9 Knockout Cell Line Successfully Recapitulate the Molecular Features of LCA4 Disease
Previous Article in Special Issue
Nuclear α-Synuclein-Derived Cytotoxic Effect via Altered Ribosomal RNA Processing in Primary Mouse Embryonic Fibroblasts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Alpha Synuclein: Neurodegeneration and Inflammation

by
Gianluigi Forloni
Department of Neuroscience, Istituto di Ricerche Farmacologiche Mario Negri IRCCS, Via Mario Negri 2, 20156 Milano, Italy
Int. J. Mol. Sci. 2023, 24(6), 5914; https://doi.org/10.3390/ijms24065914
Submission received: 10 February 2023 / Revised: 14 March 2023 / Accepted: 15 March 2023 / Published: 21 March 2023
(This article belongs to the Special Issue Neurobiology of Protein Synuclein)

Abstract

:
Alpha-Synuclein (α-Syn) is one of the most important molecules involved in the pathogenesis of Parkinson’s disease and related disorders, synucleinopathies, but also in several other neurodegenerative disorders with a more elusive role. This review analyzes the activities of α-Syn, in different conformational states, monomeric, oligomeric and fibrils, in relation to neuronal dysfunction. The neuronal damage induced by α-Syn in various conformers will be analyzed in relation to its capacity to spread the intracellular aggregation seeds with a prion-like mechanism. In view of the prominent role of inflammation in virtually all neurodegenerative disorders, the activity of α-Syn will also be illustrated considering its influence on glial reactivity. We and others have described the interaction between general inflammation and cerebral dysfunctional activity of α-Syn. Differences in microglia and astrocyte activation have also been observed when in vivo the presence of α-Syn oligomers has been combined with a lasting peripheral inflammatory effect. The reactivity of microglia was amplified, while astrocytes were damaged by the double stimulus, opening new perspectives for the control of inflammation in synucleinopathies. Starting from our studies in experimental models, we extended the perspective to find useful pointers to orient future research and potential therapeutic strategies in neurodegenerative disorders.

1. Introduction

Parkinson’s disease (PD) is the second most import neurodegenerative disorder after Alzheimer’s disease (AD), causing progressive movement disability, bradykinesia, tremor, and rigidity, accompanied by non-motor symptoms. The neuropathological hallmarks of PD are the loss of dopaminergic neurons in the substantia nigra pars compacta (SNpc) and the presence of intracellular inclusions termed Lewy bodies (LBs) to honor the neuropathologist who first described these cytoplasmic eosinophilic fibrillary bodies. In the PD brain, LBs and Lewy neurites (LN) were observed in deep mesencephalic nuclei, in SNpc, and the olfactory bulb, spreading to other regions with a prion-like mechanism in patients staged by Braak et al. (2003) [1], and later confirmed for PD patients with early onset and long duration disease, and for Lewy body dementia (LBD). In LBD, there were large amounts of LBs and LN in the cortex, and in multiple system atrophy (MSA) where LBs have also been found in glial cells, mainly oligodendroglia (glial cytoplasmic inclusions, GCI) [2].
LBs were immunocytochemically identified with anti-ubiquitin antibodies [3], but a breakthrough came when molecular genetic investigations demonstrated that autosomal dominant forms of PD were associated with mutations of the gene encoding α-synuclein (α-Syn) [4,5]. Originally sequenced as a precursor of the non-β amyloid component of senile plaque in AD (NACP) [6,7], α-Syn was identified as the main component of LB [8,9], particularly its phosphorylated form at serine residue 129 (p-Syn) [10,11]. Thus, PD and the disorders characterized by LB were collectively named synucleinopathies.
In this first quarter of the century, the biology of α-Syn in physiological conditions and its role in neurodegenerative disorders has been extensively investigated. Alpha-Syn defined natural unfolded protein is localized presynaptically, and its sequence of 140 amino acids can be structurally divided into three domains, the amphipathic N-terminal region with residues 1–60; the central non-Aβ-component (NAC) portion with residues 61–95, which is highly hydrophobic and fibrillogenic, and the acidic C-terminal region with residues 96–140. We have described the distinct functions of these three regions, including a neuroprotective activity restricted to the C-terminal region [12]. In physiological conditions, α-Syn has been described in alpha helical tetramer structures resistant to aggregation [13,14,15]. It has been suggested that tetramers undergo destabilization of their helically folded conformation prior to α-Syn aggregation into abnormal oligomeric and fibrillar assemblies [16,17]. Although the importance of the tetrameric structure of α-Syn in the physiological condition is debated in favor of the more common unstructured monomeric form [18], there is evidence that this structure is sensitive to the presence of various mutations associated with PD [19,20]. Numerous investigations have connected the physiological activity of α-Syn in monomeric forms to synaptic functionality [21,22], although there are still no conclusive results about its physiological function, while the pathological events associated with α-Syn are almost exclusively linked to its polymeric forms. Alpha-Syn can bind to lipid membrane, and adopts highly ordered α-helical conformations on lipid binding; the biophysical properties of this interaction have been extensively characterized and associated with α-Syn aggregation mechanisms [23,24,25].
The identification of α-Syn in cerebrospinal fluid (CSF) [26] and the propagation of LB pathologies from transplant to host cerebral tissue in humans [27,28] substantially changed the view on synucleinopathy etiology, and although the accumulation of α-Syn is intracellular, they became similar to the other central amyloidoses. Thus α-Syn oligomers (α-SynO) are responsible for pathological spreading and neuronal dysfunction through two independent mechanisms, as shown in prion diseases, where propagation of prion is distinct from the neurotoxic activity [29]. Furthermore, as discussed below, α-SynO are also important in the neuroinflammation associated with selective neurodegeneration in the early phase of PD and related disorders [30]. In various experimental conditions, it has been reported that α-SynO in combination with neurotoxic activity can induce glial activation [31]. On the other hand, general inflammation can influence the biological activity of α-SynO with positive or detrimental consequences for the neurotoxic processes [32,33]. The crosstalk between α-Syn and the immune system involving peripheral and central compartments is complex and fundamental in the pathogenesis of synucleinopathies [34,35,36] (Figure 1).

2. Alpha-Synuclein Oligomers

The self-assembling capacity of amyloidogenic peptides follows the nucleation mechanism proposed by Lansbury’s group thirty years ago [37]. The initial aggregation, seeding, involved a few molecules that change conformation to β sheet, from soluble oligomers to protofibrillar and fibrillar insoluble structures. Besides the common tendency to shift from random coil to β-sheet structure, the size and characteristics of monomeric proteins in neurodegenerative diseases vary widely, and a “one-dimensional crystallization” model with minimal variations [37,38] is still quite capable of describing the chemico-physical mechanism of the formation of insoluble aggregates.
The seeding mechanism responsible for protein deposition can be seen in diverse cellular compartments, with no substantial differences, and numerous factors including pH, ionic strength, temperature, mutations in amino acid sequence, metals and concentration can affect the amyloidogenesis. The pathological form of α-Syn assembles in prion-like polymeric structures by inducing the conformational conversion of native α-Syn. The process of α-Syn polymerization through oligomeric to insoluble fibrillar structures is complex, and the intermediate conformations are heterogeneous.
The recent development of cryo-EM and image processing techniques has permitted ultrastructural analysis with a resolution close to the atomic scale. These innovative techniques have been employed in various experimental conditions to analyze the ultrastructural differences in α-Syn fibrils, revealing interesting fundamental aspects of the polymorphic capacity of α-Syn aggregates. Cryo-EM analysis of recombinant α-Syn fibrils in cell-free conditions confirms the structural heterogeneity of wild-type and mutated sequences [39,40,41,42]. The analysis of fibrils from protein misfolding cyclic amplification starting from PD or MSA tissue homogenate with recombinant α-Syn showed differences from those generated only from recombinant seeds [43]. A recent cryo-EM study on α-Syn filaments from the brains of individuals with PD, PD with dementia and DLB showed them to be a single protofilament (Lewy fold), markedly different from the protofilaments of MSA [44]. Moreover, the fibrils purified from the brain present different structures from those generated in vitro using recombinant α-Syn seeded with previously extracted fibrils [44,45].
These findings confirm the structural polymorphism in relation to disease PD/LBD vs. MSA, but it is important, on the other hand, to note that multiple structures found by assembling recombinant α-Syn in vitro did not show similar heterogeneity in the analysis of brain extracted filaments. More results are needed to clarify the extent to which the structure of the fibrils formed in vitro reproduces the structural characteristics of in vivo-formed fibrils, but it appears that the nature of α-Syn filaments is less variable than expected, and similar conclusions are shared by the cryo-EM analysis of other amyloid structures [46,47,48].
The central NAC portion is the highly amyloidogenic part of α-Syn, driving the conformational change from random coil to β-sheet structure to fibrils [49]. Oligomeric structures are heterogeneous and metastable; their progression to mature fibrils is not inevitable, and alternative pathways to oligomer conformers have been described [50]. Alpha-Syn undergoes spontaneous assembly, and its aggregation capacity is influenced by mutations associated with familial PD [51,52], leading to the formation of annular protofibrils resembling a class of pore-forming bacterial toxin. Inappropriate membrane permeabilization, therefore, could be responsible for cell dysfunction and cell death [53,54] (Figure 2).
The heterogeneity of α-SynOs, influencing their toxicity, arises during their assembly. There is compelling evidence of their pathogenic role in synucleinopathies compared with monomers or fibrils that can adopt different morphologies such as spherical, chain-like, annular (pore-like structure), and tubular [17,55]. In addition to sequence mutations, α-Syn aggregation is also influenced by various post-translational modifications including phosphorylation, ubiquitination, nitration, and oxidation, which have been identified in experimental models as well as brains from PD patients. The phosphorylation at codon 129 has been involved in PD pathogenesis, but recent investigations show that this phosphorylation occurs when the aggregation is well established, and even inhibits the assembling capacity of α-Syn in the initial aggregation steps [56]. Acetylation at the N-terminal does not influence, or in some circumstances reduces, the aggregation capacity of the α-Syn full-length sequence [57,58,59].
The studies on ubiquitination are contradictory: in some cases, the self-assembling capacity of α-Syn was reduced, and in others, enhanced [58]. It has also been shown that ubiquitination may favor the α-SynO removal by proteasome [60]. More consistent are the results on the effect of glycation, which enhances the aggregation and toxicity in vitro [61] and promotes oligomerization [62]. SUMOrylation is another important post-translational modification of α-Syn that affects its accumulation, aggregation and toxicity [63].
We used the term oligomeropathies to stress the fundamental role of the oligomeric forms in the pathogenesis of virtually all protein misfolding neurodegenerative disorders; although the proteins involved in the diseases are different, the propensity to aggregate is what substantially activates a common mechanism of polymerization [38,64,65]. The nucleation of α-Syn in a structure progressively enriched with the β-sheet component is based on the initial self-assembly of a few monomers that create the oligomeric structures, heterogeneous by definition [55,66].
The affinity for membrane structures has been shown for AβOs and other oligomeric species, and α-SynOs can affect the membrane integrity by insertion in lipid bilayers [67]. However, this interaction accounts only partially for the α-SynOs’ neurotoxic effects [66]. The α-SynO oligomerization process is preferentially located on the membrane surface, and the mitochondrial membranes are more vulnerable to permeabilization [23,68,69]. This can be easily associated with the particular vulnerability of mitochondria in PD pathogenesis. Thus, α-SynOs became a potential therapeutic target in synucleinopathies [70,71].

2.1. Oligomers and Transmission of LB Pathology

In formalin-fixed paraffin-embedded brain sections from patients with PD, a novel technique to detect α-SynOs has been used to examine the distribution of α-SynOs [72]. The authors observed a difference between the distribution of α-SynOs and Lewy pathologies in PD. Alpha-SynOs were more widely distributed than Lewy pathologies, suggesting that α-SynOs may be found throughout the brain sooner in the disease course than can be detected with immunohistochemistry for Lewy pathologies.
Given the toxicity of α-SynOs, clinicopathological studies focusing on α-SynOs may give information on PD pathogenesis. These results suggest that α-SynOs may be widely distributed in PD early in the disease, and they may contribute to cognitive impairment in PD [73]. The presence α-SynOs was confirmed in the hippocampus of patients with cognitive impairment.
This important observation found direct support from our results in transgenic mice carrying the A53T human mutated sequence of α-Syn. The mutation is associated with early-onset PD, and in transgenic mice it induces cognitive decline and an accumulation of α-SynO oligomers in the hippocampus and cortex [74]. This evidence, apparently conflicting with a PD general model where the cognitive decline is secondary to motor impairment, is consistent with recent findings supporting the pathogenic role of α-SynO oligomers in PD [75]. This study used different forms of α-Syn aggregates assembled in vitro or purified from PD brain material and showed that soluble non-fibrillar α-Syn aggregates shorter than 200 nm were highly toxic species, causing lipid membrane permeabilization and inflammation. Furthermore, the hippocampus contained a higher proportion of smaller aggregates than the visual association cortex, and these hippocampal aggregates caused an increased inflammatory response [75].
It has been proposed that PD pathology can spread with a prion-like mechanism through the olfactory and vagal systems to the substantia nigra. The presence of α-Syn deposits in the olfactory bulb and anterior olfactory nucleus at Braak stage I supports this [76]. Similar evidence was reported in LBD [77]. In experimental studies, the early appearance of α-Syn deposits in the olfactory bulb has been described in transgenic mice PD models [78,79,80]. In these models, the neuropathological alterations were associated with hyposmia that appeared before the other behavioral alterations [81]. Injections of mouse or human α-Syn fibrils into the olfactory bulb of wild-type mice recruited endogenous α-Syn into pathological aggregates that spread transneuronally to over 40 other brain regions and subregions over the course of 12 months [82,83]. Inoculation of α-Syn preformed fibrils (PFF) in the olfactory bulb of non-human primates (NHP) induced a spreading of α-Syn pathology in several brain regions including SNpc and the dorsal nucleus of vagus [84].
As mentioned above, the intracellular localization of LBs and LNs in synucleinopathies implies that their formation is the consequence of a cell-to-cell passage of α-Syn. Thus, neuronal dysfunction resulting from exposure to neurotoxic oligomeric species is combined with the seeding mechanism activated by fibrils and oligomers [85,86]. Danzer et al. (2007) [87], using different protocols, prepared various α-SynO conformers and compared their biological activities in terms of cell death, cytosol calcium levels and intracellular α-Syn accumulation. In SHSY-5Y cells and primary cultures, smaller, annular structures affected calcium influx and cell death, while larger structures exerted no neurotoxic activity but promoted intracellular α-Syn aggregation, as described by Lee et al. (2010) [88]. After α-SynO exposure, the intracellular formation of large aggregates was associated with a reduction in cytoplasmic α-Syn staining, while exposure to the fibrillar form of α-Syn showed no detectable seeding effect [89]. The effect was dose- and time-dependent, and the authors demonstrated that small oligomeric structures can be converted to larger formations, activating a seeding mechanism [88].
In cultured cells overexpressing α-Syn, intracellular aggregation has also been triggered by the introduction of recombinant α-Syn fibrils incorporated in liposomes [90]. Alpha-Syn fibrils recruited endogenous soluble α-Syn proteins and converted them to insoluble, hyperphosphorylated, and ubiquitinated pathological species [90]. The exposure to α-Syn short amyloid fibrils induced intracellular aggregation in SHSY-5Y cells [89]. Alpha-SynO is secreted from neuronal cells by nonconventional exocytosis involving extracellular exosomes [91] and internalized through an active process associated or not with exosomes through an endocytic pathway [86,92,93,94]. The aggregated forms of α-Syn are associated with exosomes externally on the membrane surface or inserted within the membrane; however, internal accumulation also has been observed [86,95,96] (Figure 2).
These findings indicate that the spreading of α-Syn pathology from a cell-to-cell activating seeding intracellular mechanism is mainly associated with α-Syn in the oligomer conformation, which might be incorporated through different biological mechanisms and secreted from neuronal cells (Figure 2). In several experimental conditions, the intracellular cellular aggregation of α-Syn by exposure to PFF has also been shown [97,98,99].

2.2. Oligomer Neurotoxicity

The possibility that α-Syn might exert direct neurotoxic activity was immediately considered when the combination of genetic and neuropathological findings suggested a central role of this protein in the PD pathogenesis [100]. Since the initial studies in 2000, the possibility has been suggested that the oligomeric form of α-Syn is the main cause of neuronal dysfunction [101]. Using NAC peptide associated with a TAT sequence to vehicle the peptide inside the cells, we demonstrated selective dopaminergic toxicity in primary mesencephalic cultures and by intranigral injections in vivo [102]. Similarly to the previous investigations with Aβ peptides, α-SynO conformers were investigated in various conditions [103,104,105,106] and the possibility that different species had distinct biological activities was tested [87]. As mentioned before, the intracellular accumulation of α-Syn in pathological conditions made it harder to immediately understand the sequence of pathological events associated with α-Syn going from monomeric to insoluble fibrillar structures. The recent demonstrations that toxic oligomers can be released from fibrillar structures in cell lines and human iPSC model [66,107] and that oligomers can induce axonal damage and direct mitochondrial alterations [108] are important achievements that help to clarify the pathological process in synucleinopathies.
Recently, it has been shown that the activity of tyrosine hydroxylase, an enzyme specifically expressed in catecholamine-containing neurons, can induce the hydroxylation of Tyr 136 in the α-Syn sequence. This DOPAnization promoted the α-Syn oligomerization, which may play a pathogenic role in dopaminergic neurodegeneration [109]. The Y136-DOPA-Syn signal was detected at the striatonigral axon terminals of aged A53T Tg mice and patients with PD or MSA but not in controls. This suggests that DOPAnization of α-Syn may occur during the disease progression rather than initiating the pathological process, but in any case, it might contribute to the selective loss of dopaminergic neurons in synucleinopathies [109].
The neuronal dysfunction induced by exogenously applied α-SynOs has been investigated in various experimental conditions. We have developed a model with intracerebroventricular (ICV) injection of α-SynOs followed by behavioral, biochemical and histopathological examinations. In similar conditions, Martin et al. (2011) [110] described cognitive deficits and LTP inhibition induced by α-SynOs involving calcineurin. According to Fortuna et al. (2017) [111], a single ICV injection of α-SynOs recapitulated some of the PD pathological features, including dopaminergic cell loss and olfactory dysfunction together with motor behavior alterations. The basic pathological PD features have also been described at the experimental level with intranigral injections of α-SynOs [112], though in some conditions, α-Syn fibrils too can induce neurotoxicity [113,114,115,116].
In an experimental model developed to study the biological activities of Aβ oligomers [117,118], we compared the effects of α-Syn in monomeric, oligomeric and fibrillar forms injected ICV in wild-type mice, followed by behavioral and histopathological examinations. In contrast with the monomeric and fibrillar forms, which were ineffective, α-SynOs induced cognitive decline, seen in the object recognition test (ORT); the effect was reversible and specifically inhibited by anti- α-Syn antibody [31] and was associated with hippocampal glial activation, as observed in PD brain [75].
The neurotoxic activity of oligomers is the consequence of various interactions with cell function. Together with non-specific interferences with cell membrane that can alter its chemico-physical structures, affecting its microfluidity [119,120,121,122], two more specific mechanisms have been observed [123]. Oligomers, including α-SynOs, can directly interact with entities activating specific intracellular pathways in different conditions, although the functional consequences of this receptor interaction were not always confirmed [74,118]. Finally, several oligomeric structures formed membrane pores with ion channel function [17,87,124,125,126].
Figure 2 summarizes the possible neurotoxic mechanisms associated with exposure to α-SynOs. In neurons, α-SynOs induce perturbation of cellular and mitochondrial membranes, altering their structures and functions [71,105,127,128,129,130]. Specific receptor interactions have been proposed to explain α-synO toxicity, including prion protein [131,132], β spectrin [133,134], RLP1 receptor [135] and purinergic P2X7R receptor [136,137]. Interaction with the GluN2A NMDA receptor subunit caused synaptic dysfunction in striatum [138]. As discussed below, an essential contribution to α-synO toxicity comes from glial activation and the production of inflammatory factors affecting neuronal function (Figure 1). In this pathological scenario, the activation of TLR-2 by α-SynOs is particularly important.

3. Alpha-Synuclein and Inflammation

3.1. Peripheral and Central Inflammation

Inflammation influences the pathogenesis of virtually all neurodegenerative disorders; distinct inflammatory pathways are activated, and the timing can be different, but local or/and general inflammation is invariably involved, starting from the early phase of diseases [139]. In PD and the other synucleinopathies, dysfunction of the central and peripheral immune systems significantly participates in the development of the diseases [140].
In relation to inflammation, the increasing importance of two-way communication between the gut and the brain in neurodegenerative disorders has been supported by consistent results and numerous theoretical suggestions that await proof from experimental data [141]. However, in PD, the peripheral component of the disease is well established. PD can be considered a multi-system disorder with neuroinflammation and immune dysfunction implicated in the development of various non-motor symptoms, such as sleep and gastrointestinal dysfunction, which can precede the disease diagnosis by decades [142,143,144]. Disturbances in the gut–brain axis have been found to be associated with PD; neuropathological changes in the enteric nervous system (ENS) and significant alterations of the gut microbiota suggest an important role of these systems [145,146]. Alpha-Syn appears to have a fundamental role in the deleterious crosstalk between neuronal and immune system dysfunction in the early stages of PD pathology involving proinflammatory factors and the gut microbiota.
In the early staging of PD, α-Syn pathologies in the olfactory nucleus and vagal nerve were identified [1,147], then α-Syn inclusions in the myenteric and submucosal plexus of the ENS of postmortem PD patients [148]. This proves that α-Syn pathogenesis might originate in the gut, then propagate to the CNS.
The possibility that PD may begin with α-Syn aggregation in the ENS and then propagate to the CNS via the vagus nerve has been suggested by numerous findings [144]. In a first experiment, brain lysates from PD patients containing different types of α-Syn aggregates or synthetic PFFs were injected into the gastrointestinal wall of rats, and the transfer of aggregated α-Syn from the gut to the brain along the vagal nerve was observed [149].
However, direct induction of α-Syn pathology in the ENS by α-Syn viral overexpression or PFF injections in rodents and non-human primates (NHP) induced local pathology that did not spread outside the ENS. Rats and NHP received targeted enteric injections of PFF or adeno-associated virus overexpressing the A53T α-Syn mutated sequence associated with PD. Histopathological examination at different times after the injections showed α-Syn transient pathology in the brainstem (dorsal motor nucleus of the vagus and locus coeruleus) one month later; however, no pathology was observed at later times. Similarly, the examination of brains from injected NHP revealed no pathology despite the robust α-Syn accumulation throughout the ENS that persisted for the entire study (12 months) [150]. The authors, therefore, suggested that sustained spread of α-Syn pathology from the periphery to the CNS and subsequent propagation is rare, and that the presence of enteric α-Syn pathology and dysfunction may be an epiphenomenon [150]. This was supported by an investigation in peripheral autoptic material from PD and LBD subjects compared to age-matched controls. The staining of p-Syn, a sensitive marker of α-Syn pathologies, was seen in the vagus in 46% of LBD subjects and 89% of PD subjects, and 17% and 81%, respectively, in the stomach; no p-Syn was found in controls. The authors concluded that the involvement of peripherical organs did not distinguish all PD/LBD subjects from controls, and consequently, synucleinopathies must first be in the brain [151]. A more rational interpretation is that in the majority of DLB cases and less than 20% of PD cases, α-Syn pathology propagation is exclusively in the CNS, but although the peripheral involvement is partial, we could not rule out a peripheral start of the synucleinopathies, coexisting with cases originating inside the brain [152].
In another NHP model, either intrastriatal or enteric injection of LB extracts from PD patients induced nigrostriatal lesions and ENS pathology [153]. These findings indicated the two-way long-distance propagation of α-Syn pathology between the CNS and the ENS in PD, varying between patients and disease subtypes. However, α-Syn pathological lesions were not found in the vagal nerve, excluding this route of transmission of PD pathology.
These results suggest a possible systemic mechanism in which the general circulation acts as a route for long-distance two-way transmission of endogenous α-Syn between the enteric system and the CNS [153]. Brain extract from PD subjects was injected in this case into either the brain or ENS, while the previous investigation used recombinant α-Syn or adenovirus to induce the mutated sequence of α-Syn. We cannot rule out that other components in the cerebral material might favor the transmission of the pathology. However, in rodent models, α-Syn PFFs also propagated α-Syn pathology from the periphery to the brain. Kim et al. (2019) [154] injected α-Syn PFFs into the muscularis layer of the duodenum and pylorus in mice, showing gut-to-brain transmission of α-Syn pathology. Van Den Berge et al. (2019) [155], injecting PFFs into the wall of the duodenum in transgenic rats overexpressing human α-Syn, reported wide distribution of α-Syn pathology, including the brainstem.
It is important to stress that in these experimental studies, α-Syn transmission and deposition were constantly associated with inflammation induced by α-Syn, and in a vicious circle, inflammation promoted the propagation of α-Syn pathology [156,157,158], and intestinal inflammation increased the cerebral α-Syn accumulation [159].

3.2. Inflammation and α-Syn oligomers

The glial activation induced by ICV injection of α-SynOs was determined biochemically and immunohistopathologically, measuring the expression of IBA-1 and GFAP as markers of astrocytes and microglia, respectively, in hippocampus and cortex [31,36] (Figure 1).
In neurodegenerative disorders, the activation of TLR-2 and TLR-4 influenced the pathogenic process negatively [160,161,162,163,164,165]. We distinguished a role of TLR-4 in the AβOs and TLR-2 in the toxic effect of α-SynOs by measuring cognitive decline found similar results when glial activation was considered. The TLR-2 antagonist inhibited the glial activation induced by α-SynOs, while the TLR-4 antagonist or TLR-4 silencing abolished the glial reaction induced by AβO [17,31].
The NLRP3 inflammasome, an important component of the innate immune system, is a multimeric protein complex that assembles in response to homeostasis-altering molecular patterns (HAMPs) and certain other cellular insults [166]. The stimulation of TLRs is part of the activation of the nucleotide-binding oligomerization domain-leucine-rich repeat-pyrin domain-containing 3 (NLRP-3) inflammasome [167]. Alpha-Syn aggregates induced the secretion of proinflammatory cytokines such as IL1β and IL18 through the activation of inflammasomes, particularly the NLRP-3 [168,169]. Systemic NLRP3 inflammasome expression and the increase in IL-1β levels in the serum negatively correlate with PD progression [170].
Oral treatment with an NLRP3 inhibitor, MCC950, in wild-type mice injected with fibrillar α-Syn reduced the inflammatory factors but also the accumulation of α-Syn aggregates and dopaminergic damage [171].
Microglia are the primary innate immune cells in the brain, with an important role in maintaining cerebral homeostasis by detecting changes in their surroundings, removing cellular debris, and providing neurotrophins [172]. However, when microglia are activated by pattern recognition receptors (PRRs), they can trigger an inflammatory response that can lead to chronic neuroinflammation and neuronal damage. This chronic neuroinflammatory response is observed in patients with synucleinopathies, where microglia activation occurs in all brain regions accumulating α-Syn. Alpha-Syn can induce excessive microglial activation and inflammatory activity, which can contribute to the progression of neurodegenerative diseases [173] (Figure 1). The accumulation of α-Syn induces the secretion of IL-1β by microglia cells through activation of the NLRP3 inflammasome. Inhibition of this inflammasome has been found to prevent α-Syn pathology and dopaminergic neurodegeneration in mouse models of PD [171].
The manipulation of inflammasome at various levels improved the overall clearance of α-SynOs [173,174]. This effect was initially restricted to the activation induced by α-Syn in fibrillar form [175] but the recent results by Scheiblich et al. (2021) [163] extended the capacity to activate microglial inflammasome at the α-SynOs. The inflammasome activation by α-SynOs in microglial cells has been associated with inhibition of an autophagic mechanism and increased TNF-α release [176,177,178]. The activation of autophagy, on the other hand, affects the accumulation of α-Syn [179] overexpression of the Bcl2-associated athanogene (BAG)3 promoted autophagy and reduced the activation of the NLRP3 inflammasome induced by LPS regulating autophagy [180].
Gut inflammation may be an important driver of the systemic immune response in PD. Activation of NLRP3 inflammasome in PD has also been related to peripheral inflammation, and the microbiota–gut–brain axis has been implicated via enteric bacterial regulation of this inflammasome [181]. The levels of α-TNF, γ-interferon, IL-6 and IL-1β negatively correlated with disease duration [182].

3.3. Role of Astrocytes

The influence of general inflammation on the biological activity of α-SynOs in CNS was directly investigated in an experimental model in wild-type mice where a lasting level of inflammation was induced by a single peripheral injection of lipopolysaccharide (LPS). In comparison with control one month later, the effects of ICV injection of a sub-active dose of α-SynOs were tested by histological and behavioral examinations. The combination of LPS and a sub-active dose of α-SynOs induced cognitive decline and amplified microglial activation in the hippocampal and cortical regions, whereas astroglial reactivity was reduced with the double-hit compared to the single treatment [36]. Different behavior of astrocytes and microglia was also observed in α-Syn A53T transgenic mice injected with LPS, where microglial activation was exacerbated, while astrocyte activation was abolished.
These findings clearly illustrate the relationship between systemic inflammation and central action of α-SynOs. In our model, LPS peripheral treatment induced glial reactivity per se; the synergic effect can be restricted to the CNS. As reported by Garcia-Revilla et al. (2022) [183] in several experimental models including rotenone and 6OHDA injections, the influence of general inflammation on central neurotoxic activity is well established. The different reactivity of astrocytes and microglia is important evidence that needs to be considered in the PD pathogenesis; the astrocyte dysfunction might contribute to the neurodegenerative process in synucleinopathies [184,185].
As shown by Kuter et al. (2018) [186], prolonged astrocyte dysfunction combined with microglia activation accelerated neuronal degeneration. iPSC-derived astrocytes from PD patients with G2019S-LRRK2 mutation showed morphological atrophy and mitochondrial malfunction compared to those from age-matched control donors [187].
Brandebura et al. (2023) [188] have extensively analyzed the role of astrocytes in neurodegeneration. They underlined the changes in astrocytes at different stages of disease progression and the heterogeneity of their responses in different brain regions, distinct from any single neurodegenerative disorder. The contribution of astrocytes to neurodegeneration could be summarized in two directions: their dysfunction, which alters the physiological homeostasis indispensable for correct cerebral function, and their reactivity, which induces the production of harmful factors. Thus, alterations to astrocytes encompass both toxic gain of function and impairment of essential function. Our data fit well in this scenario, considering the astrocyte dysfunction induced by the combination of general inflammation and α-SynOs as the first part of the pathogenic mechanism in PD.

4. Therapeutic Perspectives

4.1. Synuclein Aggregation as a Pharmacological Target

Treatments to control the central and peripheral symptoms of PD are efficient and are divided to follow the progression of the disease. The numerous pharmacological and non-pharmacological tools achieve an adequate level of care especially in the early phase of the disease [189]. However, similarly for other neurodegenerative disorders, we still lack a disease-modifying drug or treatment to interfere with the pathogenesis, arresting or slowing its progression. The identification of α-Syn as the main component of LB led to experimental models being developed to demonstrate the neurotoxic activity of α-Syn in a mutated or wild-type sequence [102,190,191], and cerebral accumulation of α-Syn was immediately considered as a pharmacological target in PD and other synucleinopathies. Direct anti-amyloidogenic activity against α-Syn aggregation or modulation through the HSP70/90 pathway was initially proposed as a therapeutic approach [192,193,194,195]. Small molecules, peptides, and peptidomimetics directly targeting α-Syn and the aggregation pathways were investigated [196]. Low-molecular-weight compounds with different chemical structures have been screened in vitro, measuring their anti-aggregation activity against α-Syn’s self-assembling capacity. In the study by Masuda et al. (2006) [197], 79 compounds belonging to 12 different chemical classes inhibited α-Syn filament formation in vitro: these included polyphenols, phenothiazines, porphyrins, polyene macrolides, and Congo red and its derivatives. As expected, several of these were also able to inhibit the self-assembling aggregation of tau and β amyloid.
Other high-throughput screening methods followed to identify anti-Syn aggregation compounds [198,199], including some recent ones where known [200] or innovative methods [201,202] were applied. Numerous potential candidates to interfere with synucleinopathy pathogenesis were identified, but only a few have reached the clinical stage [203]. These include anle138b, a lead compound from a systematic high-throughput screening test combined with medicinal chemistry optimization, which showed anti-oligomerization activity against prion and α-Syn aggregates [204]. The intimate mechanism of interaction of anle138b with α-Syn fibrils has been investigated with cryo-EM, revealing stable polar interactions between anle138b and backbone moieties inside the tubular cavity of the fibrils [205]. In different PD mouse models, anle138b strongly inhibited oligomer accumulation, neuronal degeneration, and disease progression in vivo [204,206,207] and motor deficits in MSA models [208,209,210]. These findings, the optimal blood–brain barrier (BBB) passage and the oral administration strongly supported clinical investigations to test the efficacy of anle138b in PD. Two phase 1 randomized trials are evaluating the safety, tolerability and pharmacokinetics of anle13, and recruitment of the subjects has been completed in both trials. The first reported favorable safety, tolerability and pharmacokinetic parameters [211], and the second phase 1 study will establish the best drug dosage.
Squalamine is another interesting molecule that affects α-Syn aggregation, displacing α-Syn from lipid membranes, altering the first steps in its aggregation [212,213,214]. In an experimental PD model with A53T α-Syn transgenic mice, oral squalamine restored gut motility [215]. A derivative of squalamine, ENT-01, gave positive results on constipation in PD patients [216], and the results of a phase 1 clinical trial to evaluate the efficacy in PD dementia are under analysis [203]. Another aminosterol, trodusquemine, has been tested at the preclinical level to investigate its mechanism of action in inhibiting α-Syn aggregation and its efficacy in animal models [217,218] In artificial membrane, the drug is homogeneously distributed, influencing the lipid raft organization in neuronal membranes and interfering with the start of α-SynO formation [219].
The inhibition of α-Syn aggregation by peptidergic compounds is based on the synthesis of amino acid sequences targeted on the NAC core, from residues 68 to 78 of α-Syn, which bind seed fibrils to prevent their growth and elongation [220,221]. A similar approach was proposed in the past with β-sheet breakers in AD, targeting the β amyloid sequence, with interesting results in experimental models [222]. These were replicated in a prion disease model [223] but have not been translated in clinical studies.
Several other peptidergic sequences inhibited α-Syn aggregation and antagonized cell toxicity [224,225,226]. However, moving from the promising peptide sequences to an efficacious therapeutic tool is complex, and several biological obstacles must be removed, including the actual bio-availability and the BBB passage.

4.2. Immunotherapy

The immunotherapy in the AD model in the pioneer study by Schenk et al. (1999) [227], used active immunization with an Aβ sequence that reduced cerebral amyloid plaques and antagonized the cognitive behavior deficits [228].
After two decades and numerous failures, immunotherapy is now starting to be considered in AD [229]. Lecanemab, a new-generation anti-Aβ antibody, has been recently approved by the FDA for early AD treatment. Active immunization in synucleinopathies was proposed in animal models a few years after the initial experiments in an AD model [230], and passive immunization was tested later [231]. In both cases, the treatment relieved the histopathological and behavioral consequences induced by the human transgene.
Active immunization with AFFITOPE PD1A (PD01A), a new-generation immunogen, has been tested in various PD models, with positive results in terms of reduction of α-Syn deposition, protection of dopaminergic neurons and improvement of motor behavior alterations [232]. These findings have led to several phase 1 clinical trials with PD01A to identify the dosage regimen for an adequate immunological reaction and to minimize side effects in PD [233] and MSA patients [234]. Another molecule, PD03A, designed with the same technology, has also given positive results, confirmed in PD [235]. Though PD03A vaccination generated α-Syn-reactive IgG in 58% of treated patients, the titers were lower than with PD01A, where the responder rate was 89%. Clinical trials to assess the efficacy of PD01A or PD03A in synucleinopathies have not yet started.
Another synthetic peptide-based vaccine, UB-312, which targets a 12-amino-acid sequence in the C terminus of α-Syn, has been investigated in a phase 1 clinical trial with positive results [236]. The immunization with UB-312 in a Thy1SNCA/15 mouse model of PD prevented functional deficits and both central and peripheral pathology, including a reduction of α-SynO in several brain regions [237,238].
A passive immunization approach using anti-α-Syn antibodies to prevent aggregation and cell-to-cell spread of the disease has also been developed for clinical investigation. Prasinezumab is a humanized antibody that recognizes the C-terminal domain of α-Syn, based on the murine antibody 9 × 104 amply characterized in preclinical studies [231] In a phase 1 trial, prasinezumab was well tolerated and significantly reduced serum α-Syn [239]. Another trial replicated safety results but with no reduction of α-Syn in CSF [240]. The efficacy of prasinezumab was recently tested in a clinical trial in PD subjects, but the drug had no substantial effect on global or imaging measures of PD progression compared with placebo, though there was some reaction with i.v. infusion [241].
With a procedure similar to that for the purified anti-Aβ antibody aducanumab to treat AD [242], cimpanemab was obtained from isolated B-cell lines from healthy individuals. using direct ELISA [231]. The antibody recognized the N-terminal domain of the α-Syn sequence, the epitope within the initial 15 amino acids. Preclinical investigations with cimpanemab indicated some protective effects against α-Syn PFFs injected in α-Syn A53T transgenic mice [243]. A phase 1, single ascending-dose study showed that cimpanemab was well tolerated, although the CSF drug level was less than 1% of the level in plasma [244]. However, a recent phase 2 trial missed its primary and secondary endpoints, did not achieve proof of concept, and the drug was, therefore, dropped from further development [245].
MEDI1341 is another interesting anti-α-Syn antibody recently characterized in preclinical investigations, which showed a picomolar affinity for both monomeric and aggregated forms of α-Syn; in vitro and in vivo models MED1341 reduced α-Syn propagation [246]. A couple of phase 1 trials with MEDI1341 are ongoing in controls and PD subjects.
Several other anti-α-Syn antibodies are in preclinical characterization [247] but, as discussed by Jensen et al. (2023) [248], immunotherapy in synucleinopathies is not a simple issue, and before we see real positive results in clinical trials, several aspects need to be addressed. As discussed in relation to AD [249] there are various reasons for failure, including not only the complexity of the disease only partially known, but also the correct identification of the target, the timing of treatment, selection of patients, appropriate biomarkers, and correct trial design. These reasons are common to any other treatments in neurodegenerative disorders, but immunotherapy is also limited by the scarce passage of the antibodies into the brain (often, the levels of antibodies in CSF are less than 1% of those in serum) and their stability [250]. In addition, an obvious difficulty is solving a complex disease such as a neurodegenerative disorder with a single medication; combination therapy focusing on different biological targets has more chances of success.

4.3. Anti-Inflammatory Treatments

The important role of inflammation in synucleinopathies offers the opportunity to consider the control of immunoreactivity and glial activation as pharmacological targets and to develop combined therapeutic strategies focused on α-SynOs [251].
As mentioned before, ICV α-SynOs induced cognitive decline with glial activation, both effects being antagonized by TLR-2 antagonist [31]. A key role has been suggested for TLR-2 in synucleinopathies, in the CNS and periphery associated with various biological activities [173,252,253,254]. This includes a role of α-SynOs as endogenous agonist of TLR-2 on microglial cells [255] with high affinity [256]. The inhibition of autophagy by TLR-2 activation through regulation of the AKT/mTOR pathway is another mechanism acting on α-SynO accumulation, reducing the clearance of α-SynOs [257].
The different mechanisms involving α-SynOs influenced by TLR-2 include microglia and astrocytic activation, neuronal autophagy, and α-Syn transmission from neuron to neuron or from neurons to glial cells [258]. In iPSCs, the activation of TLR-2 acutely impaired the autophagy lysosomal pathway and markedly potentiated α-Syn pathology seeded with α-Syn PFF, while a novel small molecule antagonist of TLR2 (NPT1220–321) strongly reduced the effect [259]. Kouli et al. (2019) [260] found that in synucleinopathies, the effect of TLR-2 restricted to the CNS was associated with an important peripheral role of TLR-4. Xia et al. (2021) [261] reported that plasma exosomes from PD patients injected intrastriatally induced microglial activation and the propagation of exosomal α-syn from microglia to neurons; this effect appeared to be mediated by TLR-2 in microglia. These findings are encouraging the immunotherapeutic approach to PD with molecules that interfere with TLR-2.
The ICV injection of α-SynOs induced parallel neuronal dysfunction and glial activation in hippocampus, whereas systemic treatment with classical anti-inflammatory drugs such as indomethacin or ibuprofen completely abolished the glial activation, as expected, but also antagonized the memory deficits induced by oligomers injected in wild-type mice [31]. This is consistent with the fact that the well-known inflammatory agent LPS per se induces memory impairment [262]. Thus, glial activation induced by α-SynOs seemed essential for the detrimental effect on cognitive performance, and could be involved with TLR-2 at the neuronal but also the glial level, as seen above [263]. Our model simplifies the pathogenic mechanism in PD and LBD and the role of α-SynOs. However, a recent histopathological investigation in humans by Sekiya et al. (2022) [73] showed accumulation of α-SynOs in the hippocampus of PD subjects with memory impairment but also an inflammatory reaction.
We recently investigated the potential therapeutic properties of doxycycline (Doxy) in human α-syn A53T transgenic mice. Doxy is a second-generation semisynthetic tetracycline with several other pharmacological properties besides its anti-microbic capacity. These activities have led to the drug being repurposed to treat several other pathological conditions [262,263,264,265].
The combination of anti-inflammatory and anti-amyloidogenic activity turned out to counteract behavioral, biochemical and histopathological alterations in α-syn A53T transgenic mice [266]. The mice treated with Doxy were rescued from cognitive and daily activity deficits, and although no changes were observed in α-Syn deposits in the SNpc, the drug substantially reduced hippocampal and cortical α-SynO accumulation and glial activation and exerted a neuroprotective effect, overcoming LTP alterations. It is interesting that in an APP/PS1 transgenic mouse model of AD as well, Doxy’s positive effect on cognitive behavior was not immediately associated with a reduction in Aβ deposits when the drug was administered acutely or chronically for a relatively short time; the Aβ plaques were reduced only when treatment was prolonged to three months [267]. Apparently, α-SynOs are Doxy’s main target, and the action on fibrillar deposits becomes evident only when the lack of fuel has slowly reduced their size.
Our studies in different pathological contexts indicate that Doxy can directly interact with oligomers, interfering with fibrillogenesis [268]. González-Lizárraga et al. (2017) [269] showed that Doxy can modify the structure of α-SynOs into off-pathway, high-molecular-weight species; this conformer did not progress to fibrils and was not neurotoxic. This important result might explain Doxy‘s efficacy in synucleinopathies [270,271].
In a recent study by Dos Santos Pereira et al. (2022) [272] Doxy affected L-DOPA-induced dyskinesia (LID) in mice with dopaminergic impairment induced by intrastriatal injection of 6OHDA. In this model, Doxy given before L-Dopa significantly attenuated LID, and the effect was associated with a reduction in inflammatory markers in the striatum.
According to the concept of a multifactorial approach to neurodegenerative disorders, Doxy combines two different pharmacological activities in a single molecule: the inhibition of α-syn aggregation and the control of inflammation [273]. The drug is ready for a clinical challenge [256]. Its safety profile, confirmed by recent chronic treatment [274,275] and favorable pharmacokinetics [276] make Doxy an excellent candidate for PD subjects in the early phase or even in the prodromal phase.
Other compounds also have the multi-target activities of Doxy, including the anti-oligomeric activity against α-Syn0s [277,278]. Rifampicin is a well-known antibiotic with interesting pharmacological activities independently from the original anti-infectivity action. Rifampicin interacts with oligomers of different origins, including α-Syn0s, as demonstrated in vitro and in vivo [279,280]. This was associated with anti-inflammatory activities in PD and LDB models [281,282,283,284]. The anti-inflammatory activity of rifampicin in a neurological context has been shown by Bi et al. (2011) [285] using microglia BV2 cells alone or in combination with neurons to examine the neuroprotective consequences against the microglial reaction.
More recently, in zebrafish injected with rotenone, the attenuation of mitochondrial dysfunction by rifampicin was associated with a reduction in inflammatory reactivity. A similar effect has been described in human microglial cells stimulated by rotenone: in this case, the control of inflammation depended on activation of the V-type proton ATPase subunit isoform 1 (Atp6v0a1,) a subunit involved in lysosomal mechanisms [268]. In LPS-induced cognitive and motor impairment in mice, the protective effect of rifampicin appears to be mediated by the TLR4/MyD88/NF-κB signaling pathway. In synucleinopathies and the other neurodegenerative disorders, despite abundant experimental results, the clinical efficacy of rifampicin has been investigated little, with contradictory results [17,286,287,288].

5. Conclusions

The central role of α-Syn in PD and associated disorders is well established; genetic and neuropathological findings indicate that α-Syn aggregation is an initial causal event in the early phase of the disease. Alpha-Syn aggregation follows the nucleation (or seeding) model from monomer through oligomers to insoluble fibrils, common to other protein-misfolding diseases. The intracellular compartmentalization of the LB does not hamper the spread of LB pathologies through the brain and from outside (ENS) to the CNS. Both propagation from the periphery to the CNS and in the opposite direction are compatible with the development of PD, and numerous experimental studies, though not all, have proved that fibrillar α-Syn or PD brain-purified material, injected in the ENS or CNS, can spread in both directions. Numerous experimental findings point to α-SynOs, the soluble form of α-Syn aggregate, as the main agent of the pathological spread and cell-to-cell transmission, although exosomal transport can also involve other forms of aggregates [289].
Alpha-SynOs have the direct neurotoxic ability to induce neuronal dysfunction, with mechanisms shared by the other pathological oligomers [17,126]. Several findings indicate that cell-to-cell spread, intracellular seeding and neurotoxicity are distinct mechanisms. Neuropathological examinations of PD brain show clear differences between the distribution of α-SynOs and LB pathologies [73]. Together with α-Syn deposits in the early phases of synucleinopathies, an excessive inflammatory reaction sustains the pathology. Our preclinical models show that general inflammation can enhance the α-SynOs’ cerebral toxicity and that inflammation promotes the propagation of [290]. Alpha-Syn aggregation and propagation, together with the control of inflammation, are potential targets for disease-modifying treatments in PD and related disorders, which so far have been equipped only with symptomatic tools.
As generally happens with neurodegenerative diseases, numerous approaches to treat PD or MSA, including anti-amyloidogenic, small molecules or immunotherapy, have given promising results in preclinical studies but have not translated successfully to human pathological conditions. Multiple reasons explain these failures, starting from the disease complexity, the timing of treatment, selection of patients and identification of the appropriate therapeutic target and tools. All of these aspects and others in the treatment of synucleinopathies can be refined [248]. Specific efforts should focus on treatment timing: the formation of α-syn deposits and inflammation are both early events in the disease, and intervention in a prodromal phase is recommended. Another fundamental consideration involves multifunctional treatments combining diverse pharmacological targets in a single molecule or using separate compounds. Doxy is a good example to test this strategy; solid findings support its dual anti-inflammatory and anti-amyloidogenic functions with the advantages of repurposing a drug already well-known in clinical practice.
Growing knowledge about the biology of α-Syn and inflammation in neuropathological conditions continuously offers opportunities to develop new therapeutic strategies and earlier diagnostic tools [291], though they all need to be tested faster at the clinical level. The scientific community, regulatory systems and drug companies are invited to share knowledge and skills to improve the clinical protocols and speed up the evaluation of candidate treatments to interfere with the pathological causes of synucleinopathies and the other neurodegenerative disorders.

Funding

The study has been partially supported by Lugli Donation.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Braak, H.; Del Tredici, K.; Rüb, U.; de Vos, R.A.; Jansen Steur, E.N.; Braak, E. Staging of brain pathology related to sporadic Parkinson’s disease. Neurobiol. Aging 2003, 24, 197–211. [Google Scholar] [CrossRef]
  2. De Nuccio, F.; Kashyrina, M.; Serinelli, F.; Laferrière, F.; Lofrumento, D.D.; De Giorgi, F.; Ichas, F. Oligodendrocytes Prune Axons Containing α-Synuclein Aggregates In Vivo: Lewy Neurites as Precursors of Glial Cytoplasmic Inclusions in Multiple System Atrophy? Biomolecules 2023, 13, 269. [Google Scholar] [CrossRef] [PubMed]
  3. Lennox, G.; Lowe, J.; Morrell, K.; Landon, M.; Mayer, R.J. Anti-ubiquitin immunocytochemistry is more sensitive than conventional techniques in the detection of diffuse Lewy body disease. J. Neurol. Neurosurg. Psychiatry 1989, 52, 67–71. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Krüger, R.; Kuhn, W.; Müller, T.; Woitalla, D.; Graeber, M.; Kösel, S.; Przuntek, H.; Epplen, J.T.; Schöls, L.; Riess, O. Ala30Pro mutation in the gene encoding alpha-synuclein in Parkinson’s disease. Nat. Genet. 1998, 18, 106–108. [Google Scholar] [CrossRef] [PubMed]
  5. Polymeropoulos, M.H.; Lavedan, C.; Leroy, E.; Ide, S.E.; Dehejia, A.; Dutra, A.; Pike, B.; Root, H.; Rubenstein, J.; Boyer, R.; et al. Mutation in the alpha-synuclein gene identified in families with Parkinson’s disease. Science 1997, 276, 2045–2047. [Google Scholar] [CrossRef] [Green Version]
  6. Ueda, K.; Fukushima, H.; Masliah, E.; Xia, Y.; Iwai, A.; Yoshimoto, M.; Otero, D.A.; Kondo, J.; Ihara, Y.; Saitoh, T. Molecular cloning of cDNA encoding an unrecognized component of amyloid in Alzheimer disease. Proc. Natl. Acad. Sci. USA 1993, 90, 11282–11286. [Google Scholar] [CrossRef] [Green Version]
  7. Yoshimoto, M.; Iwai, A.; Kang, D.; Otero, D.A.; Xia, Y.; Saitoh, T. NACP, the precursor protein of the non-amyloid beta/A4 protein (A beta) component of Alzheimer disease amyloid, binds A beta and stimulates A beta aggregation. Proc. Natl. Acad. Sci. USA 1995, 92, 9141–9145. [Google Scholar] [CrossRef] [Green Version]
  8. Spillantini, M.G.; Schmidt, M.L.; Lee, V.M.Y.; Trojanowski, J.Q.; Jakes, R.; Goedert, M. α-Synuclein in Lewy bodies. Nature 1997, 388, 839–840. [Google Scholar] [CrossRef]
  9. Spillantini, M.G.; Crowther, R.A.; Jakes, R.; Hasegawa, M.; Goedert, M. α-Synuclein in filamentous inclusions of Lewy bodies from Parkinson’s disease and dementia with lewy bodies. Proc. Natl. Acad. Sci. USA 1998, 95, 6469–6473. [Google Scholar] [CrossRef] [Green Version]
  10. Colom-Cadena, M.; Pegueroles, J.; Herrmann, A.G.; Henstridge, C.M.; Muñoz, L.; Querol-Vilaseca, M.; Martín-Paniello, C.S.; Luque-Cabecerans, J.; Clarimon, J.; Belbin, O.; et al. Synaptic phosphorylated α-synuclein in dementia with Lewy bodies. Brain 2017, 140, 3204–3214. [Google Scholar] [CrossRef] [PubMed]
  11. Fujiwara, H.; Hasegawa, M.; Dohmae, N.; Kawashima, A.; Masliah, E.; Goldberg, M.S.; Shen, J.; Takio, K.; Iwatsubo, T. alpha-Synuclein is phosphorylated in synucleinopathy lesions. Nat. Cell Biol. 2002, 4, 160–164. [Google Scholar] [CrossRef] [PubMed]
  12. Albani, D.; Peverelli, E.; Rametta, R.; Batelli, S.; Veschini, L.; Negro, A.; Forloni, G. Protective effect of TAT-delivered alpha-synuclein: Relevance of the C-terminal domain and involvement of HSP70. FASEB J. 2004, 18, 1713–1715. [Google Scholar] [CrossRef] [PubMed]
  13. Bartels, T.; Choi, J.G.; Selkoe, D.J. α-Synuclein Occurs Physiologically as a Helically Folded Tetramer that Resists Aggregation. Nature 2011, 477, 107–110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Wang, W.; Perovic, I.; Chittuluru, J.; Kaganovich, A.; Nguyen, L.T.T.; Liao, J.; Auclair, J.R.; Johnson, D.; Landeru, A.; Simorellis, A.K.; et al. A Soluble α-Synuclein Construct Forms a Dynamic Tetramer. Proc. Natl. Acad. Sci. USA 2011, 108, 17797–17802. [Google Scholar] [CrossRef] [Green Version]
  15. Xu, L.; Bhattacharya, S.; Thompson, D. On the ubiquity of helical α-synuclein tetramers. Phys. Chem. Chem. Phys. 2019, 21, 12036–12043. [Google Scholar] [CrossRef]
  16. Dettmer, U.; Newman, A.J.; von Saucken, V.E.; Bartels, T.; Selkoe, D. KTKEGV repeat motifs are key mediators of normal α-synuclein tetramerization: Their mutation causes excess monomers and neurotoxicity. Proc. Natl. Acad. Sci. USA 2015, 112, 9596–9601. [Google Scholar] [CrossRef] [Green Version]
  17. Forloni, G.; La Vitola, P.; Balducci, C. Oligomeropathies, inflammation and prion protein binding. Front. Neurosci. 2022, 16, 822420. [Google Scholar] [CrossRef]
  18. Burré, J.; Vivona, S.; Diao, J.; Sharma, M.; Brunger, A.T.; Südhof, T.C. Properties of Native Brain α-Synuclein. Nature 2013, 498, E4–E6, discussion E6–E7. [Google Scholar] [CrossRef] [Green Version]
  19. Dettmer, U.; Newman, A.J.; Soldner, F.; Luth, E.S.; Kim, N.C.; von Saucken, V.E.; Sanderson, J.B.; Jaenisch, R.; Bartels, T.; Selkoe, D. Parkinson-Causing α-Synuclein Missense Mutations Shift Native Tetramers to Monomers as a Mechanism for Disease Initiation. Nat. Commun. 2015, 6, 7314. [Google Scholar] [CrossRef] [Green Version]
  20. Fonseca-Ornelas, L.; Stricker, J.M.S.; Soriano-Cruz, S.; Weykopf, B.; Dettmer, U.; Muratore, C.R.; Scherzer, C.R.; Selkoe, D.J. Parkinson-causing mutations in LRRK2 impair the physiological tetramerization of endogenous α-synuclein in human neurons. NP J. Parkinsons. Dis. 2022, 8, 118. [Google Scholar] [CrossRef]
  21. Gao, V.; Briano, J.A.; Komer, L.E.; Burré, J. Functional and Pathological Effects of α-Synuclein on Synaptic SNARE Complexes. J. Mol. Biol. 2023, 435, 167714. [Google Scholar] [CrossRef] [PubMed]
  22. Sharma, M.; Burré, J. α-Synuclein in synaptic function and dysfunction. Trends Neurosci. 2023, 46, 153–166. [Google Scholar] [CrossRef] [PubMed]
  23. Frieg, B.; Antonschmidt, L.; Dienemann, C.; Geraets, J.A.; Najbauer, E.E.; Matthes, D.; de Groot, B.L.; Andreas, L.B.; Becker, S.; Griesinger, C.; et al. The 3D structure of lipidic fibrils of α-synuclein. Nat. Commun. 2022, 13, 6810. [Google Scholar] [CrossRef] [PubMed]
  24. Fusco, G.; De Simone, A.; Gopinath, T.; Vostrikov, V.; Vendruscolo, M.; Dobson, C.M.; Veglia, G. Direct Observation of the Three Regions in α-Synuclein That Determine Its Membrane-Bound Behaviour. Nat. Commun. 2014, 5, 3827. [Google Scholar] [CrossRef] [Green Version]
  25. Fusco, G.; Chen, S.W.; Williamson, P.T.F.; Cascella, R.; Perni, M.; Jarvis, J.A.; Cecchi, C.; Vendruscolo, M.; Chiti, F.; Cremades, N.; et al. Structural basis of membrane disruption and cellular toxicity by α-synuclein oligomers. Science 2017, 358, 1440–1443. [Google Scholar] [CrossRef] [Green Version]
  26. Borghi, R.; Marchese, R.; Negro, A.; Marinelli, L.; Forloni, G.; Zaccheo, D.; Abbruzzese, G.; Tabaton, M. Full length alpha-synuclein is present in cerebrospinal fluid from Parkinson’s disease and normal subjects. Neurosci. Lett. 2000, 287, 65–67. [Google Scholar] [CrossRef]
  27. Kordower, J.H.; Chu, Y.; Hauser, R.A.; Freeman, T.B.; Olanow, C.W. Lewy body-like pathology in long-term embryonic nigral transplants in Parkinson’s disease. Nat. Med. 2008, 14, 504–506. [Google Scholar] [CrossRef]
  28. Li, J.Y.; Englund, E.; Holton, J.L.; Soulet, D.; Hagell, P.; Lees, A.J.; Lashley, T.; Quinn, N.P.; Rehncrona, S.; Björklund, A.; et al. Lewy bodies in grafted neurons in subjects with Parkinson’s disease suggest host-to-graft disease propagation. Nat. Med. 2008, 14, 501–503. [Google Scholar] [CrossRef]
  29. Sandberg, M.K.; Al-Doujaily, H.; Sharps, B.; Clarke, A.R.; Collinge, J. Prion propagation and toxicity in vivo occur in two distinct mechanistic phases. Nature 2011, 470, 540–542. [Google Scholar] [CrossRef]
  30. Forloni, G.; Artuso, V.; La Vitola, P.; Balducci, C. Oligomeropathies and pathogenesis of Alzheimer and Parkinson’s diseases. Mov. Disord. 2016, 31, 771–781. [Google Scholar] [CrossRef]
  31. La Vitola, P.; Balducci, C.; Cerovic, M.; Santamaria, G.; Brandi, E.; Grandi, F.; Caldinelli, L.; Colombo, L.; Morgese, M.G.; Trabace, L.; et al. Alpha-synuclein oligomers impair memory through glial cell activation and via Toll-like receptor 2. Brain Behav. Immun. 2018, 69, 591–602. [Google Scholar] [CrossRef] [PubMed]
  32. Choi, I.; Zhang, Y.; Seegobin, S.P.; Pruvost, M.; Wang, Q.; Purtell, K.; Zhang, B.; Yue, Z. Microglia clear neuron-released α-synuclein via selective autophagy and prevent neurodegeneration. Nat. Commun. 2020, 11, 1386. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Su, Y.; Shi, C.; Wang, T.; Liu, C.; Yang, J.; Zhang, S.; Fan, L.; Zheng, H.; Li, X.; Luo, H.; et al. Dysregulation of peripheral monocytes and pro-inflammation of alpha-synuclein in Parkinson’s disease. J. Neurol. 2022, 269, 6386–6394. [Google Scholar] [CrossRef] [PubMed]
  34. Venezia, S.; Refolo, V.; Polissidis, A.; Stefanis, L.; Wenning, G.K.; Stefanova, N. Toll-like receptor 4 stimulation with monophosphoryl lipid A ameliorates motor deficits and nigral neurodegeneration triggered by extraneuronal α-synucleinopathy. Mol. Neurodegener. 2017, 12, 52. [Google Scholar] [CrossRef] [PubMed]
  35. Wijeyekoon, R.S.; Kronenberg-Versteeg, D.; Scott, K.M.; Hayat, S.; Kuan, W.L.; Evans, J.R.; Breen, D.P.; Cummins, G.; Jones, J.L.; Clatworthy, M.R.; et al. Peripheral innate immune and bacterial signals relate to clinical heterogeneity in Parkinson’s disease. Brain Behav. Immun. 2020, 87, 473–488. [Google Scholar] [CrossRef]
  36. La Vitola, P.; Balducci, C.; Baroni, M.; Artioli, L.; Santamaria, G.; Castiglioni, M.; Cerovic, M.; Colombo, L.; Caldinelli, L.; Pollegioni, L.; et al. Peripheral inflammation exacerbates α-synuclein toxicity and neuropathology in Parkinson’s models. Neuropathol. Appl. Neurobiol. 2021, 47, 43–60. [Google Scholar] [CrossRef]
  37. Jarrett, J.T.; Lansbury, P.T., Jr. Seeding “one-dimensional crystallization” of amyloid: A pathogenic mechanism in Alzheimer’s disease and scrapie? Cell 1993, 73, 1055–1058. [Google Scholar] [CrossRef]
  38. Giampà, M.; Amundarain, M.J.; Herrera, M.G.; Tonali, N.; Dodero, V.I. Implementing Complementary Approaches to Shape the Mechanism of α-Synuclein Oligomerization as a Model of Amyloid Aggregation. Molecules 2021, 27, 88. [Google Scholar] [CrossRef]
  39. Boyer, D.R.; Li, B.; Sun, C.; Fan, W.; Sawaya, M.R.; Jiang, L.; Eisenberg, D.S. Structures of fibrils formed by α-synuclein hereditary disease mutant H50Q reveal new polymorphs. Nat. Struct. Mol. Biol. 2019, 26, 1044–1052. [Google Scholar] [CrossRef]
  40. Guerrero-Ferreira, R.; Taylor, N.M.; Mona, D.; Ringler, P.; Lauer, M.E.; Riek, R.; Britschgi, M.; Stahlberg, H. Cryo-EM structure of alpha-synuclein fibrils. Elife 2018, 7, e36402. [Google Scholar] [CrossRef]
  41. Sun, Y.; Hou, S.; Zhao, K.; Long, H.; Liu, Z.; Gao, J.; Zhang, Y.; Su, X.D.; Li, D.; Liu, C. Cryo-EM structure of full-length α-synuclein amyloid fibril with Parkinson’s disease familial A53T mutation. Cell Res. 2020, 30, 360–362. [Google Scholar] [CrossRef] [PubMed]
  42. Zhao, K.; Li, Y.; Liu, Z.; Long, H.; Zhao, C.; Luo, F.; Sun, Y.; Tao, Y.; Su, X.D.; Li, D.; et al. Parkinson’s disease associated mutation E46K of α-synuclein triggers the formation of a distinct fibril structure. Nat. Commun. 2020, 11, 2643. [Google Scholar] [CrossRef] [PubMed]
  43. Koga, S.; Sekiya, H.; Kondru, N.; Ross, O.A.; Dickson, D.W. Neuropathology and molecular diagnosis of Synucleinopathies. Mol. Neurodegener. 2021, 16, 83. [Google Scholar] [CrossRef] [PubMed]
  44. Yang, Y.; Shi, Y.; Schweighauser, M.; Zhang, X.; Kotecha, A.; Murzin, A.G.; Garringer, H.J.; Cullinane, P.W.; Saito, Y.; Foroud, T.; et al. Structures of α-synuclein filaments from human brains with Lewy pathology. Nature 2022, 610, 791–795. [Google Scholar] [CrossRef]
  45. Lövestam, S.; Schweighauser, M.; Matsubara, T.; Murayama, S.; Tomita, T.; Ando, T.; Hasegawa, K.; Yoshida, M.; Tarutani, A.; Hasegawa, M.; et al. Seeded assembly in vitro does not replicate the structures of α-synuclein filaments from multiple system atrophy. FEBS Open Bio 2021, 11, 999–1013. [Google Scholar] [CrossRef] [PubMed]
  46. Estaun-Panzano, J.; Arotcarena, M.L.; Bezard, E. Monitoring α-synuclein aggregation. Neurobiol. Dis. 2023, 176, 105966. [Google Scholar] [CrossRef] [PubMed]
  47. Taylor, A.I.P.; Staniforth, R.A. General Principles Underpinning Amyloid Structure. Front. Neurosci. 2022, 16, 878869. [Google Scholar] [CrossRef]
  48. Zhang, W.; Falcon, B.; Murzin, A.G.; Fan, J.; Crowther, R.A.; Goedert, M.; Scheres, S.H. Heparin-induced tau filaments are polymorphic and differ from those in Alzheimer’s and Pick’s diseases. Elife 2019, 8, e43584. [Google Scholar] [CrossRef]
  49. El-Agnaf, O.M.; Irvine, G.B. Aggregation and neurotoxicity of alpha-synuclein and related peptides. Biochem. Soc. Trans. 2002, 30, 559–565. [Google Scholar] [CrossRef]
  50. Alam, P.; Bousset, L.; Melki, R.; Otzen, D.E. α-Synuclein oligomers and fibrils: A spectrum of species, a spectrum of toxicities. J. Neurochem. 2019, 150, 522–534. [Google Scholar] [CrossRef] [Green Version]
  51. Conway, K.A.; Harper, J.D.; Lansbury, P.T. Accelerated in vitro fibril formation by a mutant alpha-synuclein linked to early-onset Parkinson disease. Nat. Med. 1998, 4, 1318–1320. [Google Scholar] [CrossRef] [PubMed]
  52. Conway, K.A.; Lee, S.J.; Rochet, J.C.; Ding, T.T.; Williamson, R.E.; Lansbury, P.T., Jr. Acceleration of oligomerization, not fibrillization, is a shared property of both alpha-synuclein mutations linked to early-onset Parkinson’s disease: Implications for pathogenesis and therapy. Proc. Natl. Acad. Sci. USA 2000, 97, 571–576. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Lashuel, H.A.; Petre, B.M.; Wall, J.; Simon, M.; Nowak, R.J.; Walz, T.; Lansbury, P.T., Jr. Alpha-synuclein, especially the Parkinson’s disease-associated mutants, forms pore-like annular and tubular protofibrils. J. Mol. Biol. 2002, 322, 1089–1102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Lashuel, H.A.; Hartley, D.; Petre, B.M.; Walz, T.; Lansbury, P.T., Jr. Neurodegenerative disease: Amyloid pores from pathogenic mutations. Nature 2002, 418, 291. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Cremades, N.; Chen, S.W.; Dobson, C.M. Structural Characteristics of α-Synuclein Oligomers. Int. Rev. Cell Mol. Biol. 2017, 329, 79–143. [Google Scholar] [PubMed]
  56. Ghanem, S.S.; Majbour, N.K.; Vaikath, N.N.; Ardah, M.T.; Erskine, D.; Jensen, N.M.; Fayyad, M.; Sudhakaran, I.P.; Vasili, E.; Melachroinou, K.; et al. α-Synuclein phosphorylation at serine 129 occurs after initial protein deposition and inhibits seeded fibril formation and toxicity. Proc. Natl. Acad. Sci. USA 2022, 119, e2109617119. [Google Scholar] [CrossRef]
  57. Lyer, A.; Sidhu, A.; Subramaniam, V. How important is the N-terminal acetylation of alpha-synuclein for its function and aggregation into amyloids? Front. Neurosci. 2022, 16, 1003997. [Google Scholar]
  58. Yoo, H.; Lee, J.; Kim, B.; Moon, H.; Jeong, H.; Lee, K.; Song, W.J.; Hur, J.K.; Oh, Y. Role of post-translational modifications on the alpha-synuclein aggregation-related pathogenesis of Parkinson’s disease. BMB Rep. 2022, 55, 323–335. [Google Scholar] [CrossRef]
  59. Watson, M.D.; Lee, J.C. N-Terminal Acetylation Affects α-Synuclein Fibril Polymorphism. Biochemistry 2019, 58, 3630–3633. [Google Scholar] [CrossRef]
  60. Ye, Y.; Klenerman, D.; Finley, D. N-Terminal Ubiquitination of Amyloidogenic Proteins Triggers Removal of Their Oligomers by the Proteasome Holoenzyme. J. Mol. Biol. 2020, 432, 585–596. [Google Scholar] [CrossRef]
  61. Chen, L.; Wei, Y.; Wang, X.; He, R. Ribosylation rapidly induces alpha-synuclein to form highly cytotoxic molten globules of advanced glycation end products. PLoS ONE 2010, 5, e9052. [Google Scholar]
  62. Vicente Miranda, H.; Szego, É.M.; Oliveira, L.M.A.; Breda, C.; Darendelioglu, E.; de Oliveira, R.M.; Ferreira, D.G.; Gomes, M.A.; Rott, R.; Oliveira, M.; et al. Glycation potentiates α-synuclein-associated neurodegeneration in synucleinopathies. Brain 2017, 140, 1399–1419. [Google Scholar] [CrossRef] [Green Version]
  63. Savyon, M.; Engelender, S. SUMOylation in α-Synuclein Homeostasis and Pathology. Front. Aging Neurosci. 2020, 12, 167. [Google Scholar] [CrossRef]
  64. Cremades, N.; Cohen, S.I.; Deas, E.; Abramov, A.Y.; Chen, A.Y.; Orte, A.; Sandal, M.; Clarke, R.W.; Dunne, P.; Aprile, F.A.; et al. Direct observation of the interconversion of normal and toxic forms of α-synuclein. Cell 2012, 149, 1048–1059. [Google Scholar] [CrossRef] [Green Version]
  65. De Oliveira, G.A.P.; Silva, J.L. Alpha-synuclein stepwise aggregation reveals features of an early onset mutation in Parkinson’s disease. Commun. Biol. 2019, 2, 374. [Google Scholar] [CrossRef] [Green Version]
  66. Cascella, R.; Chen, S.W.; Bigi, A.; Camino, J.D.; Xu, C.K.; Dobson, C.M.; Chiti, F.; Cremades, N.; Cecchi, C. The release of toxic oligomers from α-synuclein fibrils induces dysfunction in neuronal cells. Nat. Commun. 2021, 12, 1814. [Google Scholar] [CrossRef]
  67. Kim, H.Y.; Cho, M.K.; Kumar, A.; Maier, E.; Siebenhaar, C.; Becker, S.; Fernandez, C.O.; Lashuel, H.A.; Benz, R.; Lange, A.; et al. Structural properties of pore-forming oligomers of α-synuclein. JACS 2009, 131, 17482–17489. [Google Scholar] [CrossRef]
  68. Choi, M.L.; Chappard, A.; Singh, B.P.; Maclachlan, C.; Rodrigues, M.; Fedotova, E.I.; Berezhnov, A.V.; De, S.; Peddie, C.J.; Athauda, D.; et al. Pathological structural conversion of α-synuclein at the mitochondria induces neuronal toxicity. Nat. Neurosci. 2022, 25, 1134–1148. [Google Scholar] [CrossRef]
  69. Stefanovic, A.N.; Stöckl, M.T.; Claessens, M.M.; Subramaniam, V. α-Synuclein oligomers distinctively permeabilize complex model membranes. FEBS J. 2014, 281, 2838–2850. [Google Scholar] [CrossRef] [Green Version]
  70. Caruana, M.; Camilleri, A.; Farrugia, M.Y.; Ghio, S.; Jakubíčková, M.; Cauchi, R.J.; Vassallo, N. Extract from the Marine Seaweed Padina pavonica Protects Mitochondrial Biomembranes from Damage by Amyloidogenic Peptides. Molecules 2021, 26, 1444. [Google Scholar] [CrossRef]
  71. Ganjam, G.K.; Bolte, K.; Matschke, L.A.; Neitemeier, S.; Dolga, A.M.; Höllerhage, M.; Höglinger, G.U.; Adamczyk, A.; Decher, N.; Oertel, W.H.; et al. Mitochondrial damage by α-synuclein causes cell death in human dopaminergic neurons. Cell Death Dis. 2019, 10, 865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Sekiya, H.; Kowa, H.; Koga, H.; Takata, M.; Satake, W.; Futamura, N.; Funakawa, I.; Jinnai, K.; Takahashi, M.; Kondo, T.; et al. Wide distribution of alpha-synuclein oligomers in multiple system atrophy brain detected by proximity ligation. Acta Neuropathol. 2019, 137, 455–466. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Sekiya, H.; Tsuji, A.; Hashimoto, Y.; Takata, M.; Koga, S.; Nishida, K.; Futamura, N.; Kawamoto, M.; Kohara, N.; Dickson, D.W.; et al. Discrepancy between distribution of alpha-synuclein oligomers and Lewy-related pathology in Parkinson’s disease. Acta Neuropathol. Commun. 2022, 10, 13. [Google Scholar] [CrossRef] [PubMed]
  74. La Vitola, P.; Beeg, M.; Balducci, C.; Santamaria, G.; Restelli, E.; Colombo, L.; Caldinelli, L.; Pollegioni, L.; Gobbi, M.; Chiesa, R.; et al. Cellular prion protein neither binds to alpha-synuclein oligomers nor mediates their detrimental effects. Brain 2019, 142, 249–254. [Google Scholar] [CrossRef]
  75. Emin, D.; Zhang, Y.P.; Lobanova, E.; Miller, A.; Li, X.; Xia, Z.; Dakin, H.; Sideris, D.I.; Lam, J.Y.L.; Ranasinghe, R.T.; et al. Small soluble α-synuclein aggregates are the toxic species in Parkinson’s disease. Nat. Commun. 2022, 13, 5512. [Google Scholar] [CrossRef]
  76. Ubeda-Bañon, I.; Saiz-Sanchez, D.; de la Rosa-Prieto, C.; Martinez-Marcos, A. α-Synuclein in the olfactory system in Parkinson’s disease: Role of neural connections on spreading pathology. Brain Struct. Funct. 2014, 219, 1513–1526. [Google Scholar] [CrossRef]
  77. Cersosimo, M.G. Propagation of alpha-synuclein pathology from the olfactory bulb: Possible role in the pathogenesis of dementia with Lewy bodies. Cell Tissue Res. 2018, 373, 233–243. [Google Scholar] [CrossRef]
  78. Marxreiter, F.; Nuber, S.; Kandasamy, M.; Klucken, J.; Aigner, R.; Burgmayer, R.; Couillard-Despres, S.; Riess, O.; Winkler, J.; Winner, B. Changes in adult olfactory bulb neurogenesis in mice expressing the A30P mutant form of alpha-synuclein. Eur. J. Neurosci. 2009, 29, 879–890. [Google Scholar] [CrossRef] [Green Version]
  79. Taguchi, T.; Ikuno, M.; Hondo, M.; Parajuli, L.K.; Taguchi, K.; Ueda, J.; Sawamura, M.; Okuda, S.; Nakanishi, E.; Hara, J.; et al. α-Synuclein BAC transgenic mice exhibit RBD-like behaviour and hyposmia: A prodromal Parkinson’s disease model. Brain 2020, 43, 249–265. [Google Scholar] [CrossRef] [Green Version]
  80. Zhang, S.; Xiao, Q.; Le, W. Olfactory dysfunction and neurotransmitter disturbance in olfactory bulb of transgenic mice expressing human A53T mutant α-synuclein. PLoS ONE 2015, 10, e0119928. [Google Scholar] [CrossRef]
  81. Uemura, N.; Ueda, J.; Yoshihara, T.; Ikuno, M.; Uemura, M.T.; Yamakado, H.; Asano, M.; Trojanowski, J.Q.; Takahashi, R. α-Synuclein Spread from Olfactory Bulb Causes Hyposmia, Anxiety, and Memory Loss in BAC-SNCA Mice. Mov. Disord. 2021, 36, 2036–2047. [Google Scholar] [CrossRef] [PubMed]
  82. Rey, N.L.; Steiner, J.A.; Maroof, N.; Luk, K.C.; Madaj, Z.; Trojanowski, J.Q.; Lee, V.M.; Brundin, P. Widespread transneuronal propagation of α-synucleinopathy triggered in olfactory bulb mimics prodromal Parkinson’s disease. J. Exp. Med. 2016, 213, 1759–1778. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Rey, N.L.; George, S.; Steiner, J.A.; Madaj, Z.; Luk, K.C.; Trojanowski, J.Q.; Lee, V.M.; Brundin, P. Spread of aggregates after olfactory bulb injection of α-synuclein fibrils is associated with early neuronal loss and is reduced long term. Acta Neuropathol. 2018, 135, 65–83. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Sawamura, M.; Onoe, H.; Tsukada, H.; Isa, K.; Yamakado, H.; Okuda, S.; Ikuno, M.; Hatanaka, Y.; Murayama, S.; Uemura, N.; et al. Lewy Body Disease Primate Model with α-Synuclein Propagation from the Olfactory Bulb. Mov. Disord. 2022, 37, 2033–2044. [Google Scholar] [CrossRef] [PubMed]
  85. Danzer, K.M.; Krebs, S.K.; Wolff, M.; Birk, G.; Hengerer, B. Seeding induced by alpha-synuclein oligomers provides evidence for spreading of alpha-synuclein pathology. J. Neurochem. 2009, 111, 192–203. [Google Scholar] [CrossRef] [PubMed]
  86. Danzer, K.M.; Kranich, L.R.; Ruf, W.P.; Cagsal-Getkin, O.; Winslow, A.R.; Zhu, L.; Vanderburg, C.R.; McLean, P.J. Exosomal cell-to-cell transmission of alpha synuclein oligomers. Mol. Neurodegener. 2012, 7, 42. [Google Scholar] [CrossRef] [Green Version]
  87. Danzer, K.M.; Haasen, D.; Karow, A.R.; Moussaud, S.; Habeck, M.; Giese, A.; Kretzschmar, H.; Hengerer, B.; Kostka, M. Different species of alpha-synuclein oligomers induce calcium influx and seeding. J. Neurosci. 2007, 27, 9220–9232. [Google Scholar] [CrossRef] [Green Version]
  88. Lee, H.J.; Choi, C.; Lee, S.J. Membrane-bound-synuclein has a high aggregation propensity and the ability to seed the aggregation of the cytosolic form. J. Biol. Chem 2002, 277, 671–678. [Google Scholar] [CrossRef] [Green Version]
  89. Aulić, S.; Le, T.T.; Moda, F.; Abounit, S.; Corvaglia, S.; Casalis, L.; Gustincich, S.; Zurzolo, C.; Tagliavini, F.; Legname, G. Defined α-synuclein prion-like molecular assemblies spreading in cell culture. BMC Neurosci. 2014, 15, 69. [Google Scholar] [CrossRef] [Green Version]
  90. Luk, K.C.; Song, C.; O’Brien, P.; Stieber, A.; Branch, J.R.; Brunden, K.R.; Trojanowski, J.Q.; Lee, V.M. Exogenous alpha-synuclein fibrils seed the formation of Lewy body-like intracellular inclusions in cultured cells. Proc. Natl. Acad. Sci. USA 2009, 106, 20051–20056. [Google Scholar] [CrossRef] [Green Version]
  91. Emmanouilidou, E.; Melachroinou, K.; Roumeliotis, T.; Garbis, S.D.; Ntzouni, M.; Margaritis, L.H.; Stefanis, L.; Vekrellis, K. Cell-produced alpha-synuclein is secreted in a calcium-dependent manner by exosomes and impacts neuronal survival. J. Neurosci. 2010, 30, 6838–6851. [Google Scholar] [CrossRef] [Green Version]
  92. Delenclos, M.; Trendafilova, T.; Mahesh, D.; Baine, A.M.; Moussaud, S.; Yan, I.K.; Patel, T.; McLean, P.J. Investigation of Endocytic Pathways for the Internalization of Exosome-Associated Oligomeric Alpha-Synuclein. Front. Neurosci. 2017, 11, 172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Shearer, L.J.; Petersen, N.O.; Woodside, M.T. Internalization of α-synuclein oligomers into SH-SY5Y cells. Biophys. J. 2021, 120, 877–885. [Google Scholar] [CrossRef] [PubMed]
  94. Zhang, S.; Eitan, E.; Wu, T.Y.; Mattson, M.P. Intercellular transfer of pathogenic α-synuclein by extracellular vesicles is induced by the lipid peroxidation product 4-hydroxynonenal. Neurobiol. Aging. 2018, 61, 52–65. [Google Scholar] [CrossRef] [PubMed]
  95. Gustafsson, G.; Lööv, C.; Persson, E.; Lázaro, D.F.; Takeda, S.; Bergström, J.; Erlandsson, A.; Sehlin, D.; Balaj, L.; György, B.; et al. Secretion and Uptake of α-Synuclein Via Extracellular Vesicles in Cultured Cells. Cell Mol. Neurobiol. 2018, 38, 1539–1550. [Google Scholar] [CrossRef] [Green Version]
  96. Zheng, H.; Xie, Z.; Zhang, X.; Mao, J.; Wang, M.; Wei, S.; Fu, Y.; Zheng, H.; He, Y.; Chen, H.; et al. Investigation of α-Synuclein Species in Plasma Exosomes and the Oligomeric and Phosphorylated α-Synuclein as Potential Peripheral Biomarker of Parkinson’s Disease. Neuroscience 2021, 469, 79–90. [Google Scholar] [CrossRef]
  97. Volpicelli-Daley, L.A.; Luk, K.C.; Lee, V.M. Addition of exogenous α-synuclein preformed fibrils to primary neuronal cultures to seed recruitment of endogenous α-synuclein to Lewy body and Lewy neurite-like aggregates. Nat. Protoc. 2014, 9, 2135–2146. [Google Scholar] [CrossRef] [Green Version]
  98. Wu, Q.; Takano, H.; Riddle, D.M.; Trojanowski, J.Q.; Coulter, D.A.; Lee, V.M. α-Synuclein (αSyn) Preformed Fibrils Induce Endogenous α-Syn Aggregation, Compromise Synaptic Activity and Enhance Synapse Loss in Cultured Excitatory Hippocampal Neurons. J. Neurosci. 2019, 39, 5080–5094. [Google Scholar] [CrossRef] [Green Version]
  99. Wu, Q.; Shaikh, M.A.; Meymand, E.S.; Zhang, B.; Luk, K.C.; Trojanowski, J.Q.; Lee, V.M. Neuronal activity modulates alpha-synuclein aggregation and spreading in organotypic brain slice cultures and in vivo. Acta Neuropathol. 2020, 140, 831–849. [Google Scholar] [CrossRef]
  100. El-Agnaf, O.M.; Jakes, R.; Curran, M.D.; Middleton, D.; Ingenito, R.; Bianchi, E.; Pessi, A.; Neill, D.; Wallace, A. Aggregates from mutant and wild-type alpha-synuclein proteins and NAC peptide induce apoptotic cell death in human neuroblastoma cells by formation of beta-sheet and amyloid-like filaments. FEBS Lett. 1998, 440, 71–75. [Google Scholar] [CrossRef] [Green Version]
  101. Lücking, C.B.; Brice, A. Alpha-synuclein and Parkinson’s disease. Cell Mol. Life Sci. 2000, 57, 1894–1908. [Google Scholar] [CrossRef] [PubMed]
  102. Forloni, G.; Bertani, I.; Calella, A.M.; Thaler, F.; Invernizzi, R. Alpha-synuclein and Parkinson’s disease: Selective neurodegenerative effect of alpha-synuclein fragment on dopaminergic neurons in vitro and in vivo. Ann. Neurol. 2000, 47, 632–640. [Google Scholar] [CrossRef] [PubMed]
  103. Brown, D.R. Oligomeric alpha-synuclein and its role in neuronal death. IUBMB Life 2010, 62, 334–339. [Google Scholar] [CrossRef] [PubMed]
  104. 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. Biochem. J. 2012, 443, 719–726. [Google Scholar] [CrossRef] [PubMed]
  105. Jiang, Z.; de Messieres, M.; Lee, J.C. Membrane remodeling by α-synuclein and effects on amyloid formation. J. Am. Chem. Soc. 2013, 135, 15970–15973. [Google Scholar] [CrossRef] [Green Version]
  106. Rockenstein, E.; Nuber, S.; Overk, C.R.; Ubhi, K.; Mante, M.; Patrick, C.; Adame, A.; Trejo-Morales, M.; Gerez, J.; Picotti, P.; et al. Accumulation of oligomer-prone α-synuclein exacerbates synaptic and neuronal degeneration in vivo. Brain 2014, 137, 1496–1513. [Google Scholar] [CrossRef]
  107. Bigi, A.; Ermini, E.; Chen, S.W.; Cascella, R.; Cecchi, C. Exploring the Release of Toxic Oligomers from α-Synuclein Fibrils with Antibodies and STED Microscopy. Life 2021, 11, 431. [Google Scholar] [CrossRef]
  108. Prots, I.; Grosch, J.; Brazdis, R.M.; Simmnacher, K.; Veber, V.; Havlicek, S.; Hannappel, C.; Krach, F.; Krumbiegel, M.; Schütz, O.; et al. α-Synuclein oligomers induce early axonal dysfunction in human iPSC-based models of synucleinopathies. Proc. Natl. Acad. Sci. USA 2018, 115, 7813–7818. [Google Scholar] [CrossRef] [Green Version]
  109. Jin, M.; Matsumoto, S.; Ayaki, T.; Yamakado, H.; Taguchi, T.; Togawa, N.; Konno, A.; Hirai, H.; Nakajima, H.; Komai, S.; et al. DOPAnization of tyrosine in α-synuclein by tyrosine hydroxylase leads to the formation of oligomers. Nat. Commun. 2022, 13, 6880. [Google Scholar] [CrossRef]
  110. Martin, Z.S.; Neugebauer, V.; Dineley, K.T.; Kayed, R.; Zhang, W.; Reese, L.C.; Taglialatela, G. α-Synuclein oligomers oppose long-term potentiation and impair memory through a calcineurin-dependent mechanism: Relevance to human synucleopathic diseases. J. Neurochem. 2012, 120, 440–452. [Google Scholar] [CrossRef]
  111. Fortuna, J.T.S.; Gralle, M.; Beckman, D.; Neves, F.S.; Diniz, L.P.; Frost, P.S.; Barros-Aragão, F.; Santos, L.E.; Gonçalves, R.A.; Romão, L.; et al. Brain infusion of α-synuclein oligomers induces motor and non-motor Parkinson’s disease-like symptoms in mice. Behav. Brain Res. 2017, 333, 150–160. [Google Scholar] [CrossRef] [PubMed]
  112. Boi, L.; Pisanu, A.; Palmas, M.F.; Fusco, G.; Carboni, E.; Casu, M.A.; Satta, V.; Scherma, M.; Janda, E.; Mocci, I.; et al. Modeling Parkinson’s Disease Neuropathology and Symptoms by Intranigral Inoculation of Preformed Human α-Synuclein Oligomers. Int. J. Mol. Sci. 2020, 21, 8535. [Google Scholar] [CrossRef] [PubMed]
  113. Guiney, S.J.; Adlard, P.A.; Lei, P.; Mawal, C.H.; Bush, A.I.; Finkelstein, D.I.; Ayton, S. Fibrillar α-synuclein toxicity depends on functional lysosomes. J. Biol. Chem. 2020, 295, 17497–17513. [Google Scholar] [CrossRef]
  114. Peelaerts, W.; Bousset, L.; Van der Perren, A.; Moskalyuk, A.; Pulizzi, R.; Giugliano, M.; Van den Haute, C.; Melki, R.; Baekelandt, V. α-Synuclein strains cause distinct synucleinopathies after local and systemic administration. Nature 2015, 522, 340–344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Tapias, V.; Hu, X.; Luk, K.C.; Sanders, L.H.; Lee, V.M.; Greenamyre, J.T. Synthetic alpha-synuclein fibrils cause mitochondrial impairment and selective dopamine neurodegeneration in part via iNOS-mediated nitric oxide production. Cell Mol Life Sci. 2017, 74, 2851–2874. [Google Scholar] [CrossRef]
  116. Zadali, R.; Hassani, V.; Rafiei, Y.; Meratan, A.A.; Mamashli, F.; Nemat-Gorgani, M. A study on the interaction of the amyloid fibrils of α-synuclein and hen egg white lysozyme with biological membranes. Biochim. Biophys. Acta Biomembr. 2022, 1864, 183776. [Google Scholar] [CrossRef]
  117. Balducci, C.; Forloni, G. In vivo application of beta amyloid oligomers: A simple tool to evaluate mechanisms of action and new therapeutic approaches. Curr. Pharm. Des. 2014, 20, 2491–2505. [Google Scholar] [CrossRef]
  118. Balducci, C.; Beeg, M.; Stravalaci, M.; Bastone, A.; Sclip, A.; Biasini, E.; Colombo, L.; Manzoni, C.; Borsello, T.; Chiesa, R.; et al. Synthetic amyloid-beta oligomers impair long-term memory independently of cellular prion protein. Proc. Natl. Acad. Sci. USA 2010, 107, 2295–2300. [Google Scholar] [CrossRef] [Green Version]
  119. Kayed, R.; Sokolov, Y.; Edmonds, B.; McIntire, T.M.; Milton, S.C.; Hall, J.E.; Glabe, C.G. Permeabilization of lipid bilayers is a common conformation-dependent activity of soluble amyloid oligomers in protein misfolding diseases. J. Biol. Chem. 2004, 279, 46363–46366. [Google Scholar] [CrossRef] [Green Version]
  120. Madhu, P.; Das, D.; Mukhopadhyay, S. Conformation-specific perturbation of membrane dynamics by structurally distinct oligomers of Alzheimer’s amyloid-β peptide. Phys. Chem. Chem. Phys. 2021, 23, 9686–9694. [Google Scholar] [CrossRef]
  121. Pirhaghi, M.; Frank, S.A.; Alam, P.; Nielsen, J.; Sereikaite, V.; Gupta, A.; Strømgaard, K.; Andreasen, M.; Sharma, D.; Saboury, A.A.; et al. A penetratin-derived peptide reduces the membrane permeabilization and cell toxicity of α-synuclein oligomers. J. Biol. Chem. 2022, 298, 102688. [Google Scholar] [CrossRef] [PubMed]
  122. Salmona, M.; Forloni, G.; Diomede, L.; Algeri, M.; De Gioia, L.; Angeretti, N.; Giaccone, G.; Tagliavini, F.; Bugiani, O. A neurotoxic and gliotrophic fragment of the prion protein increases plasma membrane microviscosity. Neurobiol. Dis. 1997, 4, 47–57. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Benilova, I.; Karran, E.; De Strooper, B. The toxic Aβ oligomer and Alzheimer’s disease: An emperor in need of clothes. Nat. Neurosci. 2012, 15, 349–357. [Google Scholar] [CrossRef] [PubMed]
  124. Bahadi, R.; Farrelly, P.V.; Kenna, B.L.; Kourie, J.I.; Tagliavini, F.; Forloni, G.; Salmona, M. Channels formed with a mutant prion protein PrP(82-146) homologous to a 7-kDa fragment in diseased brain of GSS patients. Am. J. Physiol. Cell Physiol. 2003, 285, C862–C872. [Google Scholar] [CrossRef] [Green Version]
  125. Chen, S.W.; Drakulic, S.; Deas, E.; Ouberai, M.; Aprile, F.A.; Arranz, R.; Ness, S.; Roodveldt, C.; Guilliams, T.; De-Genst, E.J.; et al. Structural characterization of toxic oligomers that are kinetically trapped during α-synuclein fibril formation. Proc. Natl. Acad. Sci. USA 2015, 112, E1994–E2003. [Google Scholar] [CrossRef] [Green Version]
  126. Stöckl, M.T.; Zijlstra, N.; Subramaniam, V. α-Synuclein oligomers: An amyloid pore? Insights into mechanisms of α-synuclein oligomer-lipid interactions. Mol. Neurobiol. 2013, 47, 613–621. [Google Scholar] [CrossRef] [PubMed]
  127. Camilleri, A.; Zarb, C.; Caruana, M.; Ostermeier, U.; Ghio, S.; Högen, T.; Schmidt, F.; Giese, A.; Vassallo, N. Mitochondrial membrane permeabilisation by amyloid aggregates and protection by polyphenols. Biochim. Biophys. Acta. 2013, 1828, 2532–2543. [Google Scholar] [CrossRef] [Green Version]
  128. 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. J. Biol. Chem. 2014, 289, 21490–21507. [Google Scholar] [CrossRef] [Green Version]
  129. Nakamura, K.; Nemani, V.M.; Azarbal, F.; Skibinski, G.; Levy, J.M.; Egami, K.; Munishkina, L.; Zhang, J.; Gardner, B.; Wakabayashi, J.; et al. Direct membrane association drives mitochondrial fission by the Parkinson disease-associated protein alpha-synuclein. J. Biol. Chem. 2011, 286, 20710–20726. [Google Scholar] [CrossRef] [Green Version]
  130. Tomagra, G.; Franchino, C.; Cesano, F.; Chiarion, G.; de Lure, A.; Carbone, E.; Calabresi, P.; Mesin, L.; Picconi, B.; Marcantoni, A.; et al. Alpha-synuclein oligomers alter the spontaneous firing discharge of cultured midbrain neurons. Front. Cell Neurosci. 2023, 17, 1078550. [Google Scholar] [CrossRef]
  131. Corbett, G.T.; Wang, Z.; Hong, W.; Colom-Cadena, M.; Rose, J.; Liao, M.; Asfaw, A.; Hall, T.C.; Ding, L.; DeSousa, A.; et al. PrP is a central player in toxicity mediated by soluble aggregates of neurodegeneration-causing proteins. Acta Neuropathol. 2020, 139, 503–526. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Ferreira, D.G.; Temido-Ferreira, M.; Vicente Miranda, H.; Batalha, V.L.; Coelho, J.E.; Szegö, É.M.; Marques-Morgado, I.; Vaz, S.H.; Rhee, J.S.; Schmitz, M.; et al. α-synuclein interacts with PrPC to induce cognitive impairment through mGluR5 and NMDAR2B. Nat. Neurosci. 2017, 20, 1569–1579. [Google Scholar] [CrossRef] [PubMed]
  133. Maor, G.; Dubreuil, R.R.; Feany, M.B. α-Synuclein promotes neuronal dysfunction and death by disrupting the binding of ankyrin to α-Spectrin. J. Neurosci. 2023, 43, 1614–1626. [Google Scholar] [CrossRef] [PubMed]
  134. Ordonez, D.G.; Lee, M.K.; Feany, M.B. α-Synuclein induces mitochondrial dysfunction through spectrin and the actin cytoskeleton. Neuron 2018, 97, 108–124.e6. [Google Scholar] [CrossRef]
  135. Chen, K.; Martens, Y.A.; Meneses, A.; Ryu, D.H.; Lu, W.; Raulin, A.C.; Li, F.; Zhao, J.; Chen, Y.; Jin, Y.; et al. LRP1 is a neuronal receptor for α-synuclein uptake and spread. Mol. Neurodegener. 2022, 17, 57. [Google Scholar] [CrossRef] [PubMed]
  136. Wilkaniec, A.; Gąssowska, M.; Czapski, G.A.; Cieślik, M.; Sulkowski, G.; Adamczyk, A. P2X7 receptor-pannexin 1 interaction mediates extracellular alpha-synuclein-induced ATP release in neuroblastoma SH-SY5Y cells. Purinergic. Signal 2017, 13, 347–361. [Google Scholar] [CrossRef] [Green Version]
  137. Wilkaniec, A.; Cieślik, M.; Murawska, E.; Babiec, L.; Gąssowska-Dobrowolska, M.; Pałasz, E.; Jęśko, H.; Adamczyk, A. P2X7 Receptor is Involved in Mitochondrial Dysfunction Induced by Extracellular Alpha Synuclein in Neuroblastoma SH-SY5Y Cells. Int. J. Mol. Sci. 2020, 21, 3959. [Google Scholar] [CrossRef]
  138. Durante, V.; de Iure, A.; Loffredo, V.; Vaikath, N.; De Risi, M.; Paciotti, S.; Quiroga-Varela, A.; Chiasserini, D.; Mellone, M.; Mazzocchetti, P.; et al. Alpha-synuclein targets GluN2A NMDA receptor subunit causing striatal synaptic dysfunction and visuospatial memory alteration. Brain 2019, 142, 1365–1385. [Google Scholar] [CrossRef] [Green Version]
  139. Forloni, G.; Balducci, C. Alzheimer’s Disease, Oligomers, and Inflammation. J. Alzheimers. Dis. 2018, 62, 1261–1276. [Google Scholar] [CrossRef] [Green Version]
  140. Terkelsen, M.H.; Klaestrup, I.H.; Hvingelby, V.; Lauritsen, J.; Pavese, N.; Romero-Ramos, M. Neuroinflammation and Immune Changes in Prodromal Parkinson’s Disease and Other Synucleinopathies. J. Parkinsons. Dis. 2022, 12 (Suppl. S1), S149–S163. [Google Scholar] [CrossRef]
  141. Cerovic, M.; Forloni, G.; Balducci, C. Neuroinflammation and the Gut Microbiota: Possible Alternative Therapeutic Targets to Counteract Alzheimer’s Disease? Front. Aging Neurosci. 2019, 11, 284. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Saleh, H.A.; Yousef, M.H.; Abdelnaser, A. The Anti-Inflammatory Properties of Phytochemicals and Their Effects on Epigenetic Mechanisms Involved in TLR4/NF-κB-Mediated Inflammation. Front. Immunol. 2021, 12, 606069. [Google Scholar] [CrossRef] [PubMed]
  143. Tansey, M.G.; Wallings, R.L.; Houser, M.C.; Herrick, M.K.; Keating, C.E.; Joers, V. Inflammation and immune dysfunction in Parkinson disease. Nat. Rev. Immunol. 2022, 22, 657–673. [Google Scholar] [CrossRef] [PubMed]
  144. Williams, G.P.; Schonhoff, A.M.; Sette, A.; Lindestam Arlehamn, C.S. Central and Peripheral Inflammation: Connecting the Immune Responses of Parkinson’s Disease. J. Parkinsons. Dis. 2022, 12, S129–S136. [Google Scholar] [CrossRef]
  145. Pant, A.; Bisht, K.S.; Aggarwal, S.; Maiti, T.K. Human gut microbiota and Parkinson’s disease. Prog. Mol. Biol. Transl. Sci. 2022, 192, 281–307. [Google Scholar] [PubMed]
  146. Ryman, S.; Vakhtin, A.A.; Richardson, S.P.; Lin, H.C. Microbiome-gut-brain dysfunction in prodromal and symptomatic Lewy body diseases. J. Neurol. 2022; Epub ahead of print. [Google Scholar] [CrossRef] [PubMed]
  147. Parkkinen, L.; Pirttilä, T.; Alafuzoff, I. Applicability of current staging/categorization of alpha-synuclein pathology and their clinical relevance. Acta Neuropathol. 2008, 115, 399–407. [Google Scholar] [CrossRef] [Green Version]
  148. Braak, H.; de Vos, R.A.; Bohl, J.; Del Tredici, K. Gastric alpha-synuclein immunoreactive inclusions in Meissner’s and Auerbach’s plexuses in cases staged for Parkinson’s disease-related brain pathology. Neurosci. Lett 2006, 396, 67–72. [Google Scholar] [CrossRef]
  149. Holmqvist, S.; Chutna, O.; Bousset, L.; Aldrin-Kirk, P.; Li, W.; Björklund, T.; Wang, Z.-Y.; Roybon, L. Direct evidence of Parkinson pathology spread from the gastrointestinal tract to the brain in rats. Acta Neuropathol. 2014, 128, 805–820. [Google Scholar] [CrossRef] [Green Version]
  150. Manfredsson, F.P.; Luk, K.C.; Benskey, M.J.; Gezer, A.; Garcia, J.; Kuhn, N.C.; Sandoval, I.M.; Patterson, J.R.; O’Mara, A.; Yonkers, R.; et al. Induction of alpha-synuclein pathology in the enteric nervous system of the rat and non-human primate results in gastrointestinal dysmotility and transient CNS pathology. Neurobiol. Dis. 2018, 112, 106–118. [Google Scholar] [CrossRef]
  151. Beach, T.G.; Adler, C.H.; Sue, L.I.; Shill, H.A.; Driver-Dunckley, E.; Mehta, S.H.; Intorcia, A.J.; Glass, M.J.; Walker, J.E.; Arce, R.; et al. Vagus Nerve and Stomach Synucleinopathy in Parkinson’s Disease, Incidental Lewy Body Disease, and Normal Elderly Subjects: Evidence Against the “Body-First” Hypothesis. J. Parkinsons. Dis. 2021, 11, 1833–1843. [Google Scholar] [CrossRef] [PubMed]
  152. Farrow, S.L.; Cooper, A.A.; O’Sullivan, J.M. Redefining the hypotheses driving Parkinson’s diseases research. NPJ Parkinsons. Dis. 2022, 8, 45. [Google Scholar] [CrossRef] [PubMed]
  153. Arotcarena, M.L.; Dovero, S.; Prigent, A.; Bourdenx, M.; Camus, S.; Porras, G.; Thiolat, M.L.; Tasselli, M.; Aubert, P.; Kruse, N.; et al. Bidirectional gut-to-brain and brain-to-gut propagation of synucleinopathy in non-human primates. Brain 2020, 143, 1462–1475. [Google Scholar] [CrossRef]
  154. Kim, S.; Kwon, S.H.; Kam, T.I.; Panicker, N.; Karuppagounder, S.S.; Lee, S.; Lee, J.H.; Kim, W.R.; Kook, M.; Foss, C.A.; et al. Transneuronal Propagation of Pathologic α-Synuclein from the Gut to the Brain Models Parkinson’s Disease. Neuron 2019, 103, 627–641. [Google Scholar] [CrossRef] [PubMed]
  155. Van Den Berge, N.; Ferreira, N.; Gram, H.; Mikkelsen, T.W.; Alstrup, A.K.O.; Casadei, N.; Tsung-Pin, P.; Riess, O.; Nyengaard, J.R.; Tamgüney, G.; et al. Evidence for bidirectional and trans-synaptic parasympathetic and sympathetic propagation of alpha-synuclein in rats. Acta Neuropathol 2019, 138, 535–550. [Google Scholar] [CrossRef] [Green Version]
  156. Lee, H.J.; Suk, J.E.; Patrick, C.; Bae, E.J.; Cho, J.H.; Rho, S.; Hwang, D.; Masliah, E.; Lee, S.J. Direct transfer of alpha-synuclein from neuron to astroglia causes inflammatory responses in synucleinopathies. J. Biol. Chem. 2010, 285, 9262–9272. [Google Scholar] [CrossRef] [Green Version]
  157. Kim, T.K.; Bae, E.J.; Jung, B.C.; Choi, M.; Shin, S.J.; Park, S.J.; Kim, J.T.; Jung, M.K.; Ulusoy, A.; Song, M.Y.; et al. Inflammation promotes synucleinopathy propagation. Exp. Mol. Med. 2022, 54, 2148–2161. [Google Scholar] [CrossRef]
  158. Sokratous, M.C.; Dardiotis, E.; Hadjigeorgiou, G.M.; Bogdanos, D.; Xiromerisiou, G. Neurodegeneration and Inflammation-An Interesting Interplay in Parkinson’s Disease. Int. J. Mol. Sci. 2020, 21, 8421. [Google Scholar]
  159. Kishimoto, Y.; Zhu, W.; Hosoda, W.; Sen, J.M.; Mattson, M.P. Chronic Mild Gut Inflammation Accelerates Brain Neuropathology and Motor Dysfunction in α-Synuclein Mutant Mice. Neuromolecular. Med. 2019, 21, 239–249. [Google Scholar] [CrossRef]
  160. Balducci, C.; Forloni, G. TLR4 in Neurodegenerative Diseases: Alzheimer’s and Parkinson’s Diseases. In The Role of Toll-Like Receptor 4 in Infectious and Non Infectious inflammation; Rossetti, C., Peri, F., Eds.; Springer: Berlin/Heidelberg, Germany, 2021; pp. 105–118. [Google Scholar]
  161. Kollmann, T.R.; Levy, O.; Montgomery, R.R.; Goriely, S. Innate immune function by Toll-like receptors: Distinct responses in newborns and the elderly. Immunity 2012, 37, 771–783. [Google Scholar] [CrossRef] [Green Version]
  162. Lehnardt, S.; Henneke, P.; Lien, E.; Kasper, D.L.; Volpe, J.J.; Bechmann, I.; Nitsch, R.; Weber, J.R.; Golenbock, D.T.; Vartanian, T. A mechanism for neurodegeneration induced by group B streptococci through activation of the TLR2/MyD88 pathway in microglia. J. Immunol. 2006, 177, 583–592. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Leitner, G.R.; Wenzel, T.J.; Marshall, N.; Gates, E.J.; Klegeris, A. Targeting toll-like receptor 4 to modulate neuroinflammation in central nervous system disorders. Expert Opin. Ther. Targets 2019, 23, 865–882. [Google Scholar] [CrossRef]
  164. Rosenberger, K.; Derkow, K.; Dembny, P.; Krüger, C.; Schott, E.; Lehnardt, S. The impact of single and pairwise Toll-like receptor activation on neuroinflammation and neurodegeneration. J. Neuroinflamm. 2014, 11, 166. [Google Scholar] [CrossRef] [Green Version]
  165. Yang, J.; Wise, L.; Fukuchi, K.I. TLR4 Cross-Talk with NLRP3 Inflammasome and Complement Signaling Pathways in Alzheimer’s Disease. Front. Immunol. 2020, 11, 724. [Google Scholar] [CrossRef] [PubMed]
  166. Soraci, L.; Gambuzza, M.E.; Biscetti, L.; Laganà, P.; Lo Russo, C.; Buda, A.; Barresi, G.; Corsonello, A.; Lattanzio, F.; Lorello, G.; et al. Toll-like receptors and NLRP3 inflammasome-dependent pathways in Parkinson’s disease: Mechanisms and therapeutic implications. J. Neurol. 2022; Epub ahead of print. [Google Scholar] [CrossRef]
  167. Ravichandran, K.A.; Heneka, M.T. Inflammasome activation in neurodegenerative diseases. Essays Biochem. 2021, 65, 885–904. [Google Scholar] [PubMed]
  168. Haque, M.E.; Akther, M.; Jakaria, M.; Kim, I.S.; Azam, S.; Choi, D.K. Targeting the microglial NLRP3 inflammasome and its role in Parkinson’s disease. Mov. Disord. 2020, 35, 20–33. [Google Scholar] [CrossRef]
  169. Wang, X.J.; Ma, M.M.; Zhou, L.B.; Jiang, X.Y.; Hao, M.M.; Teng, R.K.F.; Wu, E.; Tang, B.S.; Li, J.Y.; Teng, J.F.; et al. Autonomic ganglionic injection of α-synuclein fibrils as a model of pure autonomic failure α-synucleinopathy. Nat. Commun. 2020, 11, 934. [Google Scholar] [CrossRef] [Green Version]
  170. Williams-Gray, C.H.; Wijeyekoon, R.; Yarnall, A.J.; Lawson, R.A.; Breen, D.P.; Evans, J.R.; Cummins, G.A.; Duncan, G.W.; Khoo, T.K.; Burn, D.J.; et al. Serum immune markers and disease progression in an incident Parkinson’s disease cohort (ICICLE-PD). Mov. Disord. 2016, 31, 995–1003. [Google Scholar] [CrossRef] [Green Version]
  171. Gordon, R.; Albornoz, E.A.; Christie, D.C.; Langley, M.R.; Kumar, V.; Mantovani, S.; Robertson, A.A.B.; Butler, M.S.; Rowe, D.B.; O’Neill, L.A.; et al. Inflammasome inhibition prevents α-synuclein pathology and dopaminergic neurodegeneration in mice. Sci. Transl. Med. 2018, 10, 465. [Google Scholar] [CrossRef] [Green Version]
  172. Colonna, M.; Butovsky, O. Microglia Function in the Central Nervous System During Health and Neurodegeneration. Annu. Rev. Immunol. 2017, 35, 441–468. [Google Scholar] [CrossRef] [PubMed]
  173. Scheiblich, H.; Bousset, L.; Schwartz, S.; Griep, A.; Latz, E.; Melki, R.; Heneka, M.T. Microglial NLRP3 Inflammasome Activation upon TLR2 and TLR5 Ligation by Distinct α-Synuclein Assemblies. J Immunol. 2021, 207, 2143–2154. [Google Scholar] [CrossRef] [PubMed]
  174. Holbrook, J.A.; Jarosz-Griffiths, H.H.; Caseley, E.; Lara-Reyna, S.; Poulter, J.A.; Williams-Gray, C.H.; Peckham, D.; McDermott, M.F. Neurodegenerative Disease and the NLRP3 Inflammasome. Front. Pharmacol. 2021, 12, 643254. [Google Scholar] [CrossRef] [PubMed]
  175. Gustot, A.; Gallea, J.I.; Sarroukh, R.; Celej, M.S.; Ruysschaert, J.M.; Raussens, V. Amyloid fibrils are the molecular trigger of inflammation in Parkinson’s disease. Biochem. J. 2015, 471, 323–333. [Google Scholar] [CrossRef] [PubMed]
  176. Bliederhaeuser, C.; Grozdanov, V.; Speidel, A.; Zondler, L.; Ruf, W.P.; Bayer, H.; Kiechle, M.; Feiler, M.S.; Freischmidt, A.; Brenner, D.; et al. Age-dependent defects of alpha-synuclein oligomer uptake in microglia and monocytes. Acta Neuropathol. 2016, 131, 379–391. [Google Scholar] [CrossRef]
  177. Du, X.Y.; Xie, X.X.; Liu, R.T. The Role of α-Synuclein Oligomers in Parkinson’s Disease. Int. J. Mol. Sci. 2020, 21, 8645. [Google Scholar] [CrossRef]
  178. Lv, Q.K.; Tao, K.X.; Wang, X.B.; Yao, X.Y.; Pang, M.Z.; Liu, J.Y.; Wang, F.; Liu, C.F. Role of α-synuclein in microglia: Autophagy and phagocytosis balance neuroinflammation in Parkinson’s disease. Inflamm. Res. 2023; Epub ahead of print. [Google Scholar] [CrossRef]
  179. Cao, Y.L.; Yang, Y.P.; Mao, C.J.; Zhang, X.Q.; Wang, C.T.; Yang, J.; Lv, D.J.; Wang, F.; Hu, L.F.; Liu, C.F. A role of BAG3 in regulating SNCA/α-synuclein clearance via selective macroautophagy. Neurobiol. Aging 2017, 60, 104–115. [Google Scholar] [CrossRef]
  180. Ying, Z.M.; Lv, Q.K.; Yao, X.Y.; Dong, A.Q.; Yang, Y.P.; Cao, Y.L.; Wang, F.; Gong, A.P.; Liu, C.F. BAG3 promotes autophagy and suppresses NLRP3 inflammasome activation in Parkinson’s disease. Ann. Transl. Med. 2022, 10, 1218. [Google Scholar] [CrossRef]
  181. Pellegrini, C.; Antonioli, L.; Calderone, V.; Colucci, R.; Fornai, M.; Blandizzi, C. Microbiota-gut-brain axis in health and disease: Is NLRP3 inflammasome at the crossroads of microbiota-gut-brain communications? Prog. Neurobiol. 2020, 191, 101806. [Google Scholar] [CrossRef]
  182. Devos, D.; Lebouvier, T.; Lardeux, B.; Biraud, M.; Rouaud, T.; Pouclet, H.; Coron, E.; Bruley des Varannes, S.; Naveilhan, P.; Nguyen, J.M.; et al. Colonic inflammation in Parkinson’s disease. Neurobiol. Dis. 2013, 50, 42–48. [Google Scholar] [CrossRef] [PubMed]
  183. García-Revilla, J.; Herrera, A.J.; de Pablos, R.M.; Venero, J.L. Inflammatory Animal Models of Parkinson’s Disease. J. Parkinsons. Dis. 2022, 12 (Suppl. S1), S165–S182. [Google Scholar] [CrossRef] [PubMed]
  184. Chen, C.; Mossman, E.; Malko, P.; McDonald, D.; Blain, A.P.; Bone, L.; Erskine, D.; Filby, A.; Vincent, A.E.; Hudson, G.; et al. Astrocytic Changes in Mitochondrial Oxidative Phosphorylation Protein Levels in Parkinson’s Disease. Mov. Disord. 2022, 37, 302–314. [Google Scholar] [CrossRef] [PubMed]
  185. Wang, C.; Yang, T.; Liang, M.; Xie, J.; Song, N. Astrocyte dysfunction in Parkinson’s disease: From the perspectives of transmitted α-synuclein and genetic modulation. Transl. Neurodegener. 2021, 10, 39. [Google Scholar] [CrossRef] [PubMed]
  186. Kuter, K.; Olech, Ł.; Głowacka, U. Prolonged Dysfunction of Astrocytes and Activation of Microglia Accelerate Degeneration of Dopaminergic Neurons in the Rat Substantia Nigra and Block Compensation of Early Motor Dysfunction Induced by 6-OHDA. Mol. Neurobiol. 2018, 55, 3049–3066. [Google Scholar] [CrossRef] [Green Version]
  187. Ramos-Gonzalez, P.; Mato, S.; Chara, J.C.; Verkhratsky, A.; Matute, C.; Cavaliere, F. Astrocytic atrophy as a pathological feature of Parkinson’s disease with LRRK2 mutation. NPJ Parkinsons. Dis. 2021, 7, 31. [Google Scholar] [CrossRef]
  188. Brandebura, A.N.; Paumier, A.; Onur, T.S.; Allen, N.J. Astrocyte contribution to dysfunction, risk and progression in neurodegenerative disorders. Nat. Rev. Neurosci. 2023, 24, 23–39. [Google Scholar] [CrossRef]
  189. Jankovic, J.; Tan, E.K. Parkinson’s disease: Etiopathogenesis and treatment. J. Neurol. Neurosurg. Psychiatry 2020, 91, 795–808. [Google Scholar] [CrossRef] [PubMed]
  190. Feany, M.B.; Bender, W.W. A Drosophila model of Parkinson’s disease. Nature 2000, 404, 394–398. [Google Scholar] [CrossRef]
  191. Xu, M.; Loa-Kum-Cheung, W.; Zhang, H.; Quinn, R.J.; Mellick, G.D. Identification of a New α-Synuclein Aggregation Inhibitor via Mass Spectrometry Based Screening. ACS Chem. Neurosci. 2019, 10, 2683–2691. [Google Scholar] [CrossRef]
  192. Auluck, P.K.; Bonini, N.M. Pharmacological prevention of Parkinson disease in Drosophila. Nat. Med. 2002, 8, 1185–1186. [Google Scholar] [CrossRef] [PubMed]
  193. Auluck, P.K.; Chan, H.Y.; Trojanowski, J.Q.; Lee, V.M.; Bonini, N.M. Chaperone suppression of alpha-synuclein toxicity in a Drosophila model for Parkinson’s disease. Science 2002, 295, 865–868. [Google Scholar] [CrossRef] [PubMed]
  194. Paleologou, K.E.; Irvine, G.B.; El-Agnaf, O.M. Alpha-synuclein aggregation in neurodegenerative diseases and its inhibition as a potential therapeutic strategy. Biochem. Soc. Trans. 2005, 33, 1106–1110. [Google Scholar] [CrossRef] [PubMed]
  195. Windisch, M.; Hutter-Paier, B.; Rockenstein, E.; Hashimoto, M.; Mallory, M.; Masliah, E. Development of a new treatment for Alzheimer’s disease and Parkinson’s disease using anti-aggregatory beta-synuclein-derived peptides. J. Mol. Neurosci. 2002, 19, 63–69. [Google Scholar] [CrossRef] [PubMed]
  196. Vidović, M.; Rikalovic, M.G. Alpha-Synuclein Aggregation Pathway in Parkinson’s Disease: Current Status and Novel Therapeutic Approaches. Cells 2022, 11, 1732. [Google Scholar] [CrossRef] [PubMed]
  197. Masuda, M.; Suzuki, N.; Taniguchi, S.; Oikawa, T.; Nonaka, T.; Iwatsubo, T.; Hisanaga, S.; Goedert, M.; Hasegawa, M. Small molecule inhibitors of alpha-synuclein filament assembly. Biochemistry 2006, 45, 6085–6094. [Google Scholar] [CrossRef] [PubMed]
  198. Pujols, J.; Peña-Díaz, S.; Conde-Giménez, M.; Pinheiro, F.; Navarro, S.; Sancho, J.; Ventura, S. High-Throughput Screening Methodology to Identify Alpha-Synuclein Aggregation Inhibitors. Int. J. Mol. Sci. 2017, 18, 478. [Google Scholar] [CrossRef] [Green Version]
  199. Xu, J.; Kao, S.Y.; Lee, F.J.; Song, W.; Jin, L.W.; Yankner, B.A. Dopamine-dependent neurotoxicity of alpha-synuclein: A mechanism for selective neurodegeneration in Parkinson disease. Nat. Med. 2002, 8, 600–606. [Google Scholar] [CrossRef]
  200. Hideshima, M.; Kimura, Y.; Aguirre, C.; Kakuda, K.; Takeuchi, T.; Choong, C.J.; Doi, J.; Nabekura, K.; Yamaguchi, K.; Nakajima, K.; et al. Two-step screening method to identify α-synuclein aggregation inhibitors for Parkinson’s disease. Sci. Rep. 2022, 12, 351. [Google Scholar] [CrossRef]
  201. Casalino, E.; Stine, L.B.; Corin, A.J.; Thai, C.T.; Quiroz, J.; Wilson, S.C.; Labow, M.; Mittal, S. A novel high-throughput screening strategy for targeting alpha-synuclein and other long-lived proteins. SLAS Discov. 2022, 27, 349–357. [Google Scholar] [CrossRef]
  202. De Luca, L.; Vittorio, S.; Peña-Díaz, S.; Pitasi, G.; Fornt-Suñé, M.; Bucolo, F.; Ventura, S.; Gitto, R. Ligand-Based Discovery of a Small Molecule as Inhibitor of α-Synuclein Amyloid Formation. Int. J. Mol. Sci. 2022, 23, 14844. [Google Scholar] [CrossRef] [PubMed]
  203. Grosso Jasutkar, H.; Oh, S.E.; Mouradian, M.M. Therapeutics in the Pipeline Targeting α-Synuclein for Parkinson’s Disease. Pharmacol. Rev. 2022, 74, 207–237. [Google Scholar] [CrossRef] [PubMed]
  204. Wagner, J.; Ryazanov, S.; Leonov, A.; Levin, J.; Shi, S.; Schmidt, F.; Prix, C.; Pan-Montojo, F.; Bertsch, U.; Mitteregger-Kretzschmar, G.; et al. Anle138b: A novel oligomer modulator for disease-modifying therapy of neurodegenerative diseases such as prion and Parkinson’s disease. Acta Neuropathol. 2013, 125, 795–813. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Antonschmidt, L.; Matthes, D.; Dervişoğlu, R.; Frieg, B.; Dienemann, C.; Leonov, A.; Nimerovsky, E.; Sant, V.; Ryazanov, S.; Giese, A.; et al. The clinical drug candidate anle138b binds in a cavity of lipidic α-synuclein fibrils. Nat. Commun. 2022, 13, 5385. [Google Scholar] [CrossRef]
  206. Levin, J.; Schmidt, F.; Boehm, C.; Prix, C.; Bötzel, K.; Ryazanov, S.; Leonov, A.; Griesinger, C.; Giese, A. The oligomer modulator anle138b inhibits disease progression in a Parkinson mouse model even with treatment started after disease onset. Acta Neuropathol. 2014, 127, 779–780. [Google Scholar] [CrossRef] [Green Version]
  207. Wegrzynowicz, M.; Bar-On, D.; Calo’, L.; Anichtchik, O.; Iovino, M.; Xia, J.; Ryazanov, S.; Leonov, A.; Giese, A.; Dalley, J.W.; et al. Depopulation of dense α-synuclein aggregates is associated with rescue of dopamine neuron dysfunction and death in a new Parkinson’s disease model. Acta Neuropathol. 2019, 138, 575–595. [Google Scholar] [CrossRef] [Green Version]
  208. Fellner, L.; Kuzdas-Wood, D.; Levin, J.; Ryazanov, S.; Leonov, A.; Griesinger, C.; Giese, A.; Wenning, G.K.; Stefanova, N. Anle138b Partly Ameliorates Motor Deficits Despite Failure of Neuroprotection in a Model of Advanced Multiple System Atrophy. Front. Neurosci. 2016, 10, 99. [Google Scholar] [CrossRef] [Green Version]
  209. Heras-Garvin, A.; Weckbecker, D.; Ryazanov, S.; Leonov, A.; Griesinger, C.; Giese, A.; Wenning, G.K.; Stefanova, N. Anle138b modulates α-synuclein oligomerization and prevents motor decline and neurodegeneration in a mouse model of multiple system atrophy. Mov. Disord. 2019, 34, 255–263. [Google Scholar] [CrossRef] [Green Version]
  210. Lemos, M.; Venezia, S.; Refolo, V.; Heras-Garvin, A.; Schmidhuber, S.; Giese, A.; Leonov, A.; Ryazanov, S.; Griesinger, C.; Galabova, G.; et al. Targeting α-synuclein by PD03 AFFITOPE® and Anle138b rescues neurodegenerative pathology in a model of multiple system atrophy: Clinical relevance. Transl. Neurodegener 2020, 9, 38. [Google Scholar] [CrossRef]
  211. Levin, J.; Sing, N.; Melbourne, S.; Morgan, A.; Mariner, C.; Spillantini, M.G.; Wegrzynowicz, M.; Dalley, J.W.; Langer, S.; Ryazanov, S.; et al. Safety, tolerability and pharmacokinetics of the oligomer modulator anle138b with exposure levels sufficient for therapeutic efficacy in a murine Parkinson model: A randomised, double-blind, placebo-controlled phase 1a trial. EBioMedicine 2022, 80, 104021. [Google Scholar] [CrossRef]
  212. Dou, T.; Kurouski, D. Phosphatidylcholine and Phosphatidylserine Uniquely Modify the Secondary Structure of α-Synuclein Oligomers Formed in Their Presence at the Early Stages of Protein Aggregation. ACS Chem. Neurosci. 2022, 13, 2380–2385. [Google Scholar] [CrossRef] [PubMed]
  213. Limbocker, R.; Staats, R.; Chia, S.; Ruggeri, F.S.; Mannini, B.; Xu, C.K.; Perni, M.; Cascella, R.; Bigi, A.; Sasser, L.R.; et al. Squalamine and Its Derivatives Modulate the Aggregation of Amyloid-β and α-Synuclein and Suppress the Toxicity of Their Oligomers. Front. Neurosci. 2021, 15, 680026. [Google Scholar] [CrossRef] [PubMed]
  214. Perni, M.; Galvagnion, C.; Maltsev, A.; Meisl, G.; Müller, M.B.; Challa, P.K.; Kirkegaard, J.B.; Flagmeier, P.; Cohen, S.I.; Cascella, R.; et al. A natural product inhibits the initiation of α-synuclein aggregation and suppresses its toxicity. Proc. Natl. Acad. Sci. USA 2017, 114, E1009–E1017. [Google Scholar] [CrossRef] [Green Version]
  215. West, C.L.; Mao, Y.K.; Delungahawatta, T.; Amin, J.Y.; Farhin, S.; McQuade, R.M.; Diwakarla, S.; Pustovit, R.; Stanisz, A.M.; Bienenstock, J.; et al. Squalamine Restores the Function of the Enteric Nervous System in Mouse Models of Parkinson’s Disease. J. Parkinsons. Dis. 2020, 10, 1477–1491. [Google Scholar] [CrossRef] [PubMed]
  216. Camilleri, M.; Subramanian, T.; Pagan, F.; Isaacson, S.; Gil, R.; Hauser, R.A.; Feldman, M.; Goldstein, M.; Kumar, R.; Truong, D.; et al. Oral ENT-01 Targets Enteric Neurons to Treat Constipation in Parkinson Disease: A Randomized Controlled Trial. Ann. Intern. Med. 2022, 175, 1666–1674. [Google Scholar] [CrossRef]
  217. Limbocker, R.; Errico, S.; Barbut, D.; Knowles, T.P.J.; Vendruscolo, M.; Chiti, F.; Zasloff, M. Squalamine and trodusquemine: Two natural products for neurodegenerative diseases, from physical chemistry to the clinic. Nat. Prod. Rep. 2022, 39, 742–753. [Google Scholar] [CrossRef] [PubMed]
  218. Perni, M.; Flagmeier, P.; Limbocker, R.; Cascella, R.; Aprile, F.A.; Galvagnion, C.; Heller, G.T.; Meisl, G.; Chen, S.W.; Kumita, J.R.; et al. Multistep Inhibition of α-Synuclein Aggregation and Toxicity in Vitro and in Vivo by Trodusquemine. ACS Chem. Biol. 2018, 13, 2308–2319. [Google Scholar] [CrossRef]
  219. Barletti, B.; Lucchesi, G.; Muscat, S.; Errico, S.; Barbut, D.; Danani, A.; Zasloff, M.; Grasso, G.; Chiti, F.; Caminati, G. Reorganization of the outer layer of a model of the plasma membrane induced by a neuroprotective aminosterol. Colloids Surf. B Biointerfaces 2022, 222, 113115. [Google Scholar] [CrossRef]
  220. Meade, R.M.; Morris, K.J.; Watt, K.J.C.; Williams, R.J.; Mason, J.M. The Library Derived 4554W Peptide Inhibits Primary Nucleation of α-Synuclein. J. Mol. Biol. 2020, 432, 166706. [Google Scholar] [CrossRef]
  221. Sangwan, S.; Sahay, S.; Murray, K.A.; Morgan, S.; Guenther, E.L.; Jiang, L.; Williams, C.K.; Vinters, H.V.; Goedert, M.; Eisenberg, D.S. Inhibition of synucleinopathic seeding by rationally designed inhibitors. Elife 2020, 9, e46775. [Google Scholar] [CrossRef]
  222. Soto, C.; Sigurdsson, E.M.; Morelli, L.; Kumar, R.A.; Castaño, E.M.; Frangione, B. Beta-sheet breaker peptides inhibit fibrillogenesis in a rat brain model of amyloidosis: Implications for Alzheimer’s therapy. Nat. Med. 1998, 4, 822–826. [Google Scholar] [CrossRef] [PubMed]
  223. Soto, C.; Kascsak, R.J.; Saborío, G.P.; Aucouturier, P.; Wisniewski, T.; Prelli, F.; Kascsak, R.; Mendez, E.; Harris, D.A.; Ironside, J.; et al. Reversion of prion protein conformational changes by synthetic beta-sheet breaker peptides. Lancet 2000, 355, 192–197. [Google Scholar] [CrossRef] [PubMed]
  224. Jin, J.W.; Fan, X.; Del Cid-Pellitero, E.; Liu, X.X.; Zhou, L.; Dai, C.; Gibbs, E.; He, W.; Li, H.; Wu, X.; et al. Development of an α-synuclein knockdown peptide and evaluation of its efficacy in Parkinson’s disease models. Commun. Biol. 2021, 4, 232. [Google Scholar] [CrossRef] [PubMed]
  225. Mitra, A.; Sarkar, N. Sequence and structure-based peptides as potent amyloid inhibitors: A review. Arch. Biochem. Biophys. 2020, 695, 108614. [Google Scholar] [CrossRef]
  226. Popova, B.; Wang, D.; Rajavel, A.; Dhamotharan, K.; Lázaro, D.F.; Gerke, J.; Uhrig, J.F.; Hoppert, M.; Outeiro, T.F.; Braus, G.H. Identification of Two Novel Peptides that Inhibit α-Synuclein Toxicity and Aggregation. Front. Mol. Neurosci. 2021, 14, 659926. [Google Scholar] [CrossRef]
  227. Schenk, D.B.; Koller, M.; Ness, D.K.; Griffith, S.G.; Grundman, M.; Zago, W.; Soto, J.; Atiee, G.; Ostrowitzki, S.; Kinney, G.G. First-in-human assessment of PRX002, an anti-α-synuclein monoclonal antibody, in healthy volunteers. Mov. Disord. 2017, 32, 211–218. [Google Scholar] [CrossRef]
  228. Morgan, D.; Diamond, D.M.; Gottschall, P.E.; Ugen, K.E.; Dickey, C.; Hardy, J.; Duff, K.; Jantzen, P.; DiCarlo, G.; Wilcock, D.; et al. A beta peptide vaccination prevents memory loss in an animal model of Alzheimer’s disease. Nature 2000, 408, 982–985. [Google Scholar] [CrossRef]
  229. van Dyck, C.H.; Swanson, C.J.; Aisen, P.; Bateman, R.J.; Chen, C.; Gee, M.; Kanekiyo, M.; Li, D.; Reyderman, L.; Cohen, S.; et al. Lecanemab in Early Alzheimer’s Disease. N. Engl. J. Med. 2023, 388, 9–21. [Google Scholar] [CrossRef]
  230. Masliah, E.; Rockenstein, E.; Adame, A.; Alford, M.; Crews, L.; Hashimoto, M.; Seubert, P.; Lee, M.; Goldstein, J.; Chilcote, T.; et al. Effects of alpha-synuclein immunization in a mouse model of Parkinson’s disease. Neuron 2005, 46, 857–868. [Google Scholar] [CrossRef] [Green Version]
  231. Masliah, E.; Rockenstein, E.; Mante, M.; Crews, L.; Spencer, B.; Adame, A.; Patrick, C.; Trejo, M.; Ubhi, K.; Rohn, T.T.; et al. Passive immunization reduces behavioral and neuropathological deficits in an alpha-synuclein transgenic model of Lewy body disease. PLoS ONE 2011, 6, e19338. [Google Scholar] [CrossRef] [Green Version]
  232. Mandler, M.; Valera, E.; Rockenstein, E.; Weninger, H.; Patrick, C.; Adame, A.; Santic, R.; Meindl, S.; Vigl, B.; Smrzka, O.; et al. Next-generation active immunization approach for synucleinopathies: Implications for Parkinson’s disease clinical trials. Acta Neuropathol. 2014, 127, 861–879. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Volc, D.; Poewe, W.; Kutzelnigg, A.; Lührs, P.; Thun-Hohenstein, C.; Schneeberger, A.; Galabova, G.; Majbour, N.; Vaikath, N.; El-Agnaf, O.; et al. Safety and immunogenicity of the α-synuclein active immunotherapeutic PD01A in patients with Parkinson’s disease: A randomised, single-blinded, phase 1 trial. Lancet Neurol. 2020, 19, 591–600. [Google Scholar] [CrossRef] [PubMed]
  234. Meissner, W.G.; Traon, A.P.; Foubert-Samier, A.; Galabova, G.; Galitzky, M.; Kutzelnigg, A.; Laurens, B.; Lührs, P.; Medori, R.; Péran, P.; et al. A Phase 1 Randomized Trial of Specific Active α-Synuclein Immunotherapies PD01A and PD03A in Multiple System Atrophy. Mov. Disord. 2020, 35, 1957–1965. [Google Scholar] [CrossRef]
  235. Poewe, W.; Volc, D.; Seppi, K.; Medori, R.; Lührs, P.; Kutzelnigg, A.; Djamshidian, A.; Thun-Hohenstein, C.; Meissner, W.G.; Rascol, O.; et al. Safety and Tolerability of Active Immunotherapy Targeting α-Synuclein with PD03A in Patients with Early Parkinson’s Disease: A Randomized, Placebo-Controlled, Phase 1 Study. J. Parkinsons. Dis. 2021, 11, 1079–1089. [Google Scholar] [CrossRef] [PubMed]
  236. Yu, H.J.; Thijssen, E.; van Brummelen, E.; van der Plas, J.L.; Radanovic, I.; Moerland, M.; Hsieh, E.; Groeneveld, G.J.; Dodart, J.C. A Randomized First-in-Human Study with UB-312, a UBITh® α-Synuclein Peptide Vaccine. Mov. Disord. 2022, 37, 1416–1424. [Google Scholar] [CrossRef]
  237. Nimmo, J.T.; Verma, A.; Dodart, J.C.; Wang, C.Y.; Savistchenko, J.; Melki, R.; Carare, R.O.; Nicoll, J.A.R. Novel antibodies detect additional α-synuclein pathology in synucleinopathies: Potential development for immunotherapy. Alzheimers Res. Ther. 2020, 12, 159. [Google Scholar] [CrossRef] [PubMed]
  238. Nimmo, J.T.; Smith, H.; Wang, C.Y.; Teeling, J.L.; Nicoll, J.A.R.; Verma, A.; Dodart, J.C.; Liu, Z.; Lin, F.; Carare, R.O. Immunisation with UB-312 in the Thy1SNCA mouse prevents motor performance deficits and oligomeric α-synuclein accumulation in the brain and gut. Acta Neuropathol. 2022, 143, 55–73. [Google Scholar] [CrossRef]
  239. Schenk, D.; Barbour, R.; Dunn, W.; Gordon, G.; Grajeda, H.; Guido, T.; Hu, K.; Huang, J.; Johnson-Wood, K.; Khan, K.; et al. Immunization with amyloid-beta attenuates Alzheimer-disease-like pathology in the PDAPP mouse. Nature 1999, 400, 173–177. [Google Scholar] [CrossRef]
  240. Jankovic, J.; Goodman, I.; Safirstein, B.; Marmon, T.K.; Schenk, D.B.; Koller, M.; Zago, W.; Ness, D.K.; Griffith, S.G.; Grundman, M.; et al. Safety and Tolerability of Multiple Ascending Doses of PRX002/RG7935, an Anti-α-Synuclein Monoclonal Antibody, in Patients with Parkinson Disease: A Randomized Clinical Trial. JAMA Neurol. 2018, 75, 1206–1214. [Google Scholar] [CrossRef] [Green Version]
  241. Pagano, G.; Taylor, K.I.; Anzures-Cabrera, J.; Marchesi, M.; Simuni, T.; Marek, K.; Postuma, R.B.; Pavese, N.; Stocchi, F.; Azulay, J.P.; et al. Trial of Prasinezumab in Early-Stage Parkinson’s Disease. N. Engl. J. Med. 2022, 387, 421–432. [Google Scholar] [CrossRef]
  242. Sevigny, J.; Chiao, P.; Bussière, T.; Weinreb, P.H.; Williams, L.; Maier, M.; Dunstan, R.; Salloway, S.; Chen, T.; Ling, Y.; et al. The antibody aducanumab reduces Aβ plaques in Alzheimer’s disease. Nature 2016, 537, 50–56. [Google Scholar] [CrossRef] [PubMed]
  243. Weihofen, A.; Liu, Y.; Arndt, J.W.; Huy, C.; Quan, C.; Smith, B.A.; Baeriswyl, J.L.; Cavegn, N.; Senn, L.; Su, L.; et al. Development of an aggregate-selective, human-derived α-synuclein antibody BIIB054 that ameliorates disease phenotypes in Parkinson’s disease models. Neurobiol. Dis. 2019, 124, 276–288. [Google Scholar] [CrossRef]
  244. Brys, M.; Fanning, L.; Hung, S.; Ellenbogen, A.; Penner, N.; Yang, M.; Welch, M.; Koenig, E.; David, E.; Fox, T.; et al. Randomized phase I clinical trial of anti-α-synuclein antibody BIIB054. Mov. Disord. 2019, 34, 1154–1163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Lang, A.E.; Siderowf, A.D.; Macklin, E.A.; Poewe, W.; Brooks, D.J.; Fernandez, H.H.; Rascol, O.; Giladi, N.; Stocchi, F.; Tanner, C.M.; et al. Trial of Cinpanemab in Early Parkinson’s Disease. N. Engl. J. Med. 2022, 387, 408–420. [Google Scholar] [CrossRef] [PubMed]
  246. Schofield, D.J.; Irving, L.; Calo, L.; Bogstedt, A.; Rees, G.; Nuccitelli, A.; Narwal, R.; Petrone, M.; Roberts, J.; Brown, L.; et al. Preclinical development of a high affinity α-synuclein antibody, MEDI1341, that can enter the brain, sequester extracellular α-synuclein and attenuate α-synuclein spreading in vivo. Neurobiol. Dis. 2019, 132, 104582. [Google Scholar] [CrossRef]
  247. Brendza, R.; Gao, X.; Stark, K.L.; Lin, H.; Lee, S.H.; Hu, C.; Cai, H.; Di Cara, D.; Hsiao, Y.C.; Ngu, H.; et al. Anti-α-synuclein c-terminal antibodies block PFF uptake and accumulation of phospho-synuclein in preclinical models of Parkinson’s disease. Neurobiol. Dis. 2022, 177, 105969. [Google Scholar] [CrossRef] [PubMed]
  248. Jensen, P.H.; Schlossmacher, M.G.; Stefanis, L. Who Ever Said It Would Be Easy? Reflecting on Two Clinical Trials Targeting α-Synuclein. Mov. Disord. 2023; Epub ahead of print. [Google Scholar] [CrossRef]
  249. Forloni, G. Alzheimer’s disease: From basic science to precision medicine approach. BMJ Neurol. Open. 2020, 2, e000079. [Google Scholar] [CrossRef]
  250. Höllerhage, M.; Wolff, A.; Chakroun, T.; Evsyukov, V.; Duan, L.; Chua, O.W.; Tang, Q.; Koeglsperger, T.; Höglinger, G.U. Binding Stability of Antibody-α-Synuclein Complexes Predicts the Protective Efficacy of Anti-α-synuclein Antibodies. Mol. Neurobiol. 2022, 59, 3980–3995. [Google Scholar] [CrossRef]
  251. Nguyen, L.T.N.; Nguyen, H.D.; Kim, Y.J.; Nguyen, T.T.; Lai, T.T.; Lee, Y.K.; Ma, H.I.; Kim, Y.E. Role of NLRP3 Inflammasome in Parkinson’s Disease and Therapeutic Considerations. J. Parkinsons. Dis. 2022, 12, 2117–2133. [Google Scholar] [CrossRef]
  252. Dutta, D.; Jana, M.; Majumder, M.; Mondal, S.; Roy, A.; Pahan, K. Selective targeting of the TLR2/MyD88/NF-κB pathway reduces α-synuclein spreading in vitro and in vivo. Nat. Commun. 2021, 12, 5382. [Google Scholar] [CrossRef] [PubMed]
  253. Sun, L.; Jiang, W.W.; Wang, Y.; Yuan, Y.S.; Rong, Z.; Wu, J.; Fan, Y.; Lu, M.; Zhang, K.Z. Phosphorylated α-synuclein aggregated in Schwann cells exacerbates peripheral neuroinflammation and nerve dysfunction in Parkinson’s disease through TLR2/NF-κB pathway. Cell Death Discov. 2021, 7, 289. [Google Scholar] [CrossRef] [PubMed]
  254. Xia, X.; Li, Z.; Liu, K.; Wu, Y.; Jiang, D.; Lai, Y. Staphylococcal LTA-Induced miR-143 Inhibits Propionibacterium acnes-Mediated Inflammatory Response in Skin. J. Investig. Dermatol. 2016, 136, 621–630. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  255. Kim, C.; Ho, D.H.; Suk, J.E.; You, S.; Michael, S.; Kang, J.; Lee, S.J.; Masliah, E.; Hwang, D.; Lee, H.J.; et al. Neuron-released oligomeric α-synuclein is an endogenous agonist of TLR2 for paracrine activation of microglia. Nat. Commun. 2013, 4, 1562. [Google Scholar] [CrossRef] [Green Version]
  256. Kumari, M.; Hanpude, P.; Maiti, T.K. α-Synuclein Exhibits Differential Membrane Perturbation, Nucleation, and TLR2 Binding through Its Secondary Structure. ACS Chem. Neurosci. 2020, 11, 4203–4214. [Google Scholar] [CrossRef]
  257. Kim, C.; Rockenstein, E.; Spencer, B.; Kim, H.K.; Adame, A.; Trejo, M. Antagonizing Neuronal Toll-like Receptor 2 Prevents Synucleinopathy by Activating Autophagy. Cell Rep. 2015, 13, 771–782. [Google Scholar] [CrossRef] [Green Version]
  258. Kwon, S.; Iba, M.; Masliah, E.; Kim, C. Targeting Microglial and Neuronal Toll-like Receptor 2 in Synucleinopathies. Exp. Neurobiol. 2019, 28, 547–553. [Google Scholar] [CrossRef] [Green Version]
  259. Chedid, J.; Labrador-Garrido, A.; Zhong, S.; Gao, J.; Zhao, Y.; Perera, G.; Kim, W.S.; Halliday, G.M.; Dzamko, N. A small molecule toll-like receptor antagonist rescues α-synuclein fibril pathology. J. Biol. Chem. 2022, 298, 102260. [Google Scholar] [CrossRef]
  260. Kouli, A.; Horne, C.B.; Williams-Gray, C.H. Toll-like receptors and their therapeutic potential in Parkinson’s disease and α-synucleinopathies. Brain Behav. Imm. 2019, 81, 41–51. [Google Scholar]
  261. Xia, Y.; Zhang, G.; Kou, L.; Yin, S.; Han, C.; Hu, J.; Wan, F.; Sun, Y.; Wu, J.; Li, Y.; et al. Reactive microglia enhance the transmission of exosomal α-synuclein via toll-like receptor 2. Brain 2021, 144, 2024–2037. [Google Scholar] [CrossRef]
  262. Balducci, C.; Santamaria, G.; La Vitola, P.; Brandi, E.; Grandi, F.; Viscomi, A.R. Doxycycline counteracts neuroinflammation restoring memory in Alzheimer’s disease mouse models. Neurobiol. Agin. 2018, 70, 128–139. [Google Scholar] [CrossRef] [PubMed]
  263. Forloni, G.; Iussich, S.; Awan, T.; Colombo, L.; Angeretti, N.; Girola, L.; Bertani, I.; Poli, G.; Caramelli, M.; Bruzzone, M.; et al. Tetracyclines affect prion infectivity. Proc. Natl. Acad. Sci. USA 2002, 99, 10849–10854. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Haïk, S.; Marcon, G.; Mallet, A.; Tettamanti, M.; Welaratne, A.; Giaccone, G.; Azimi, S.; Pietrini, V.; Fabreguettes, J.R.; Imperiale, D.; et al. Doxycycline in Creutzfeldt-Jakob disease: A phase 2, randomised, double-blind, placebo-controlled trial. Lancet Neurol. 2014, 13, 150–158. [Google Scholar] [CrossRef] [PubMed]
  265. Balducci, C.; Forloni, G. Doxycycline for Alzheimer’s Disease: Fighting β-Amyloid Oligomers and Neuroinflammation. Front. Pharmacol. 2019, 10, 738. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  266. La Vitola, P.; Artioli, L.; Cerovic, M.; Poletto, C.; Dacomo, L.; Leva, S.; Balducci, C.; Forloni, G. Repositioning Doxycycline for treating Parkinson’s Disease: Evidence from a pre-clinical mouse model. Parkinsonism. Relat. Dis. 2023, 106, 105229. [Google Scholar] [CrossRef]
  267. Balducci, C.; Frasca, A.; Zotti, M.; La Vitola, P.; Mhillaj, E.; Grigoli, E.; Iacobellis, M.; Grandi, F.; Messa, M.; Colombo, L.; et al. Toll-like receptor 4-dependent glial cell activation mediates the impairment in memory establishment induced by β-amyloid oligomers in an acute mouse model of Alzheimer’s disease. Brain Behav. Immun. 2017, 60, 188–197. [Google Scholar] [CrossRef]
  268. Tagliavini, F.; Forloni, G.; Colombo, L.; Rossi, G.; Girola, L.; Canciani, B.; Angeretti, N.; Giampaolo, L.; Peressini, E.; Awan, T.; et al. Tetracycline affects abnormal properties of synthetic PrP peptides and PrP(Sc) in vitro. J. Mol. Biol. 2000, 300, 1309–1322. [Google Scholar] [CrossRef]
  269. González-Lizárraga, F.; Socías, S.B.; Ávila, C.L.; Torres-Bugeau, C.M.; Barbosa, L.R.; Binolfi, A.; Sepúlveda-Díaz, J.E.; Del-Bel, E.; Fernandez, C.O.; Papy-Garcia, D.; et al. Repurposing doxycycline for synucleinopathies: Remodelling of α-synuclein oligomers towards non-toxic parallel beta-sheet structured species. Sci. Rep. 2017, 7, 41755. [Google Scholar] [CrossRef] [Green Version]
  270. Dominguez-Meijide, A.; Parrales, V.; Vasili, E.; González-Lizárraga, F.; König, A.; Lázaro, D.F.; Lannuzel, A.; Haik, S.; Del Bel, E.; Chehín, R.; et al. Doxycycline inhibits α-synuclein-associated pathologies in vitro and in vivo. Neurobiol. Dis. 2021, 151, 105256. [Google Scholar] [CrossRef]
  271. Forloni, G.; La Vitola, P.; Cerovic, M.; Balducci, C. Inflammation and Parkinson’s disease pathogenesis: Mechanisms and therapeutic insight. Prog. Mol. Biol. Transl. Sci. 2021, 177, 175–202. [Google Scholar]
  272. Dos Santos Pereira, M.; do Nascimento, G.C.; Bortolanza, M.; Michel, P.P.; Raisman-Vozari, R.; Del Bel, E. Doxycycline attenuates l-DOPA-induced dyskinesia through an anti-inflammatory effect in a hemiparkinsonian mouse model. Front. Pharmacol. 2022, 13, 1045465. [Google Scholar] [CrossRef] [PubMed]
  273. Bortolanza, M.; Nascimento, G.C.; Socias, S.B.; Ploper, D.; Chehín, R.N.; Raisman-Vozari, R.; Del-Bel, E. Tetracycline repurposing in neurodegeneration: Focus on Parkinson’s disease. J. Neural. Transm. 2018, 125, 1403–1415. [Google Scholar] [CrossRef] [PubMed]
  274. Forloni, G.; Tettamanti, M.; Lucca, U.; Albanese, Y.; Quaglio, E.; Chiesa, R.; Villani, F.; Erbetta, E.; Redaelli, V.; Tagliavini, F.; et al. Preventive study in subjects at risk of FFI: Innovative approach to rare diseases. Prion 2015, 9, 75–79. [Google Scholar] [CrossRef] [PubMed]
  275. Forloni, G.; Roiter, I.; Artuso, V.; Marcon, M.; Colesso, W.; Luban, E.; Lucca, U.; Tettamanti, M.; Pupillo, E.; Redaelli, V.; et al. Preventive pharmacological treatment in subjects at risk for fatal familial insomnia: Science and public engagement. Prion 2022, 16, 66–77. [Google Scholar] [CrossRef]
  276. Lucchetti, J.; Fracasso, C.; Balducci, C.; Passoni, A.; Forloni, G.; Salmona, M.; Gobbi, M. Plasma and Brain Concentrations of Doxycycline after Single and Repeated Doses in Wild-Type and APP23 Mice. J. Pharmacol. Exp. Ther. 2019, 368, 32–40. [Google Scholar] [CrossRef] [Green Version]
  277. Naoi, M.; Maruyama, W.; Shamoto-Nagai, M. Disease-modifying treatment of Parkinson’s disease by phytochemicals: Targeting multiple pathogenic factors. J. Neural. Transm. 2022, 129, 737–753. [Google Scholar] [CrossRef]
  278. Zhong, J.; Li, M.; Xu, J.; Dong, W.; Qin, Y.; Qiu, S.; Li, X.; Wang, H. Roflupram attenuates α-synuclein-induced cytotoxicity and promotes the mitochondrial translocation of Parkin in SH-SY5Y cells overexpressing A53T mutant α-synuclein. Toxicol. Appl. Pharmacol. 2022, 436, 115859. [Google Scholar] [CrossRef]
  279. Umeda, T.; Hatanaka, Y.; Sakai, A.; Tomiyama, T. Nasal Rifampicin Improves Cognition in a Mouse Model of Dementia with Lewy Bodies by Reducing α-Synuclein Oligomers. Int. J. Mol. Sci. 2021, 22, 8453. [Google Scholar] [CrossRef]
  280. Yurtsever, İ.; Üstündağ, Ü.V.; Ünal, İ.; Ateş, P.S.; Emekli-Alturfan, E. Rifampicin decreases neuroinflammation to maintain mitochondrial function and calcium homeostasis in rotenone-treated zebrafish. Drug Chem. Toxicol. 2022, 45, 1544–1551. [Google Scholar] [CrossRef]
  281. Acuña, L.; Corbalán, N.S.; Raisman-Vozari, R. Rifampicin quinone pretreatment improves neuronal survival by modulating microglia inflammation induced by α-synuclein. Neural Regen. Res. 2020, 15, 1473–1474. [Google Scholar] [CrossRef]
  282. Acuña, L.; Hamadat, S.; Corbalán, N.S.; González-Lizárraga, F.; Dos-Santos-Pereira, M.; Rocca, J.; Díaz, J.S.; Del-Bel, E.; Papy-García, D.; Chehín, R.N.; et al. Rifampicin and Its Derivative Rifampicin Quinone Reduce Microglial Inflammatory Responses and Neurodegeneration Induced In Vitro by α-Synuclein Fibrillary Aggregates. Cells 2019, 8, 776. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  283. Bi, W.; Cheng, X.; Zeng, Z.; Zhou, R.; Luo, R.; Zhang, J.; Zhu, L. Rifampicin ameliorates lipopolysaccharide-induced cognitive and motor impairments via inhibition of the TLR4/MyD88/NF-κB signaling pathway in mice. Neurol. Res. 2021, 43, 358–371. [Google Scholar] [CrossRef] [PubMed]
  284. Liang, Y.; Zheng, D.; Peng, S.; Lin, D.; Jing, X.; Zeng, Z.; Chen, Y.; Huang, K.; Xie, Y.; Zhou, T.; et al. Rifampicin attenuates rotenone-treated microglia inflammation via improving lysosomal function. Toxicol. In Vitro 2020, 63, 104690. [Google Scholar] [CrossRef] [PubMed]
  285. Bi, W.; Zhu, L.; Wang, C.; Liang, Y.; Liu, J.; Shi, Q.; Tao, E. Rifampicin inhibits microglial inflammation and improves neuron survival against inflammation. Brain Res. 2011, 13, 12–20. [Google Scholar] [CrossRef] [PubMed]
  286. Eschlböck, S.; Krismer, F.; Wenning, G.K. Interventional trials in atypical parkinsonism. Parkinsonism. Relat. Disord. 2016, 22 (Suppl. S1), S82–S92. [Google Scholar] [CrossRef]
  287. Singer, W.; Low, P.A. Optimizing clinical trial design for multiple system atrophy: Lessons from the rifampicin study. Clin. Auton. Res. 2015, 25, 47–52. [Google Scholar] [CrossRef] [Green Version]
  288. Yulug, B.; Hanoglu, L.; Ozansoy, M.; Isık, D.; Kilic, U.; Kilic, E.; Schabitz, W.R. Therapeutic role of rifampicin in Alzheimer’s disease. Psychiatry Clin. Neurosci. 2018, 72, 152–159. [Google Scholar] [CrossRef] [Green Version]
  289. Brás, I.C.; Khani, M.H.; Vasili, E.; Möbius, W.; Riedel, D.; Parfentev, I.; Gerhardt, E.; Fahlbusch, C.; Urlaub, H.; Zweckstetter, M.; et al. Molecular Mechanisms Mediating the Transfer of Disease-Associated Proteins and Effects on Neuronal Activity. J. Parkinsons. Dis. 2022, 12, 2397–2422. [Google Scholar] [CrossRef]
  290. Kim, C.; Spencer, B.; Rockenstein, E.; Yamakado, H.; Mante, M.; Adame, A.; Fields, J.A.; Masliah, D.; Iba, M.; Lee, H.J.; et al. Immunotherapy targeting toll-like receptor 2 alleviates neurodegeneration in models of synucleinopathy by modulating α-synuclein transmission and neuroinflammation. Mol. Neurodegener. 2018, 13, 43. [Google Scholar] [CrossRef]
  291. Al-Qassabi, A.; Tsao, T.S.; Racolta, A.; Kremer, T.; Cañamero, M.; Belousov, A.; Santana, M.A.; Beck, R.C.; Zhang, H.; Meridew, J.; et al. Immunohistochemical Detection of Synuclein Pathology in Skin in Idiopathic Rapid Eye Movement Sleep Behavior Disorder and Parkinsonism. Mov. Disord. 2021, 36, 895–904. [Google Scholar] [CrossRef]
Figure 1. The pathological scenario in synucleinopathies. α-Syn in various forms, monomeric, oligomeric and fibrils, interact with neurons and glial cells. Neuronal dysfunction is induced by direct α-Syn oligomers, originating from outside or inside the cells. The permanent activation of microglial cells can induce the production of factors with detrimental effects on neuronal function, and astrocyte activation or dysfunction may contribute to neuronal damage.
Figure 1. The pathological scenario in synucleinopathies. α-Syn in various forms, monomeric, oligomeric and fibrils, interact with neurons and glial cells. Neuronal dysfunction is induced by direct α-Syn oligomers, originating from outside or inside the cells. The permanent activation of microglial cells can induce the production of factors with detrimental effects on neuronal function, and astrocyte activation or dysfunction may contribute to neuronal damage.
Ijms 24 05914 g001
Figure 2. α-Syn and neuronal dysfunction. The fibrillogenic process from monomers to Lewy bodies passes through soluble aggregates with α-Syn oligomer as an intermediate conformer. These soluble aggregates may interact directly with neuronal membrane through specific acceptor, pore formation or unspecific interference with double lipid structure. Endocytosis and exocytosis through exosomes may contribute to the formation of Lewy bodies and cell-to-cell spread of the pathology. The progress of α-Syn oligomers to protofibrils and fibrils is not inevitable, and in specific conditions, stable oligomers can be formed (oligomer off) with no further assembling.
Figure 2. α-Syn and neuronal dysfunction. The fibrillogenic process from monomers to Lewy bodies passes through soluble aggregates with α-Syn oligomer as an intermediate conformer. These soluble aggregates may interact directly with neuronal membrane through specific acceptor, pore formation or unspecific interference with double lipid structure. Endocytosis and exocytosis through exosomes may contribute to the formation of Lewy bodies and cell-to-cell spread of the pathology. The progress of α-Syn oligomers to protofibrils and fibrils is not inevitable, and in specific conditions, stable oligomers can be formed (oligomer off) with no further assembling.
Ijms 24 05914 g002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Forloni, G. Alpha Synuclein: Neurodegeneration and Inflammation. Int. J. Mol. Sci. 2023, 24, 5914. https://doi.org/10.3390/ijms24065914

AMA Style

Forloni G. Alpha Synuclein: Neurodegeneration and Inflammation. International Journal of Molecular Sciences. 2023; 24(6):5914. https://doi.org/10.3390/ijms24065914

Chicago/Turabian Style

Forloni, Gianluigi. 2023. "Alpha Synuclein: Neurodegeneration and Inflammation" International Journal of Molecular Sciences 24, no. 6: 5914. https://doi.org/10.3390/ijms24065914

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

Article Metrics

Back to TopTop