You are viewing a javascript disabled version of the site. Please enable Javascript for this site to function properly.
Go to headerGo to navigationGo to searchGo to contentsGo to footer
In content section. Select this link to jump to navigation

Implication of Alpha-Synuclein Phosphorylation at S129 in Synucleinopathies: What Have We Learned in the Last Decade?

Abstract

Abnormal accumulation of proteinaceous intraneuronal inclusions called Lewy bodies (LBs) is the neurpathological hallmark of Parkinson’s disease (PD) and related synucleinopathies. These inclusions are mainly constituted of a presynaptic protein, α-synuclein (α-syn). Over the past decade, growing amounts of studies reported an aberrant accumulation of phosphorylated α-syn at the residue S129 (pS129) in the brain of patients suffering from PD, as well as in transgenic animal models of synucleinopathies. Whereas only a small fraction of α-syn (<4%) is phosphorylated in healthy brains, a dramatic accumulation of pS129 (>90%) has been observed within LBs, suggesting that this post-translational modification may play an important role in the regulation of α-syn aggregation, LBs formation and neuronal degeneration. However, whether phosphorylation at S129 suppresses or enhances α-syn aggregation and toxicity in vivo remains a subject of active debate. The answer to this question has important implications for understanding the role of phosphorylation in the pathogenesis of synucleinopathies and determining if targeting kinases or phosphatases could be a viable therapeutic strategy for the treatment of these devastating neurological disorders. In the present review, we explore recent findings from in vitro, cell-based assays and in vivo studies describing the potential implications of pS129 in the regulation of α-syn physiological functions, as well as its implication in synucleinopathies pathogenesis and diagnosis.

INTRODUCTION

Parkinson’s disease (PD) and related synucleinopathies, including dementia with Lewy bodies (DLB) and multiple system atrophy (MSA), are characterized by the progressive loss of vulnerable neuronal populations in the brain [1–3] and the presence of intraneuronal α-synuclein (α-syn)-rich inclusions, called Lewy bodies (LBs) [4–7]. Converging evidence from neuropathological, in vivo and in vitro studies support a causal role of α-syn in the pathogenesis of synucleinopathies [6, 7]. However, the molecular and cellular mechanisms controlling α-syn aggregation and toxicity remain in part non-elucidated.

In the last few years, an increasing number of studies reported that α-syn within LBs is subjected to several post-translational modifications (PTMs), including phosphorylation, ubiquitination, cross-linking, truncations and nitration, suggesting that these modifications may play a key role in the regulation of α-syn aggregation and toxicity in vivo [8]. Among these PTMs, a growing interest has been focused on the phosphorylation at residue S129 (pS129) and its possible implication on α-syn-induced neurodegeneration [8, 9]. Indeed, under normal conditions, only a small fraction (>4%) of α-syn is constitutively phosphorylated at S129 in the brain [10–12], whereas a dramatic accumulation (<90%) of pS129 has been observed in the brains of patients suffering from synucleinopathies [12–15], as well as in transgenic animal models of PD [16–19]. These findings strongly support the hypothesis that phosphorylation at S129 may play an important role in the control of α-syn normal functions, as well as the regulation of its aggregation, LBs formation and neurotoxicity.

Despite increased efforts to identify the kinase responsible for α-syn phosphorylation, and to understand the consequences of such modifications on the biophysical and biochemical properties of α-syn, the questions on how phosphorylation modulates α-syn normal functions and whether pS129 suppresses or enhances α-syn toxicity in vivo remain the subjects of heated debate. The answer to this question has important implications for understanding the exact role of α-syn phosphorylation in the pathogenesis of PD and related disorders and may lead to the identification of new therapeutic targets for the treatment of these neurological disorders.

Phosphorylation is an important molecular switch for the regulation of α-syn proprieties and physiological functions

Although the exact native state of α-syn remains the subject of active investigation and debate [6, 20, 21], converging in vitro [21–23] and in vivo [21, 24, 25] data suggest that this protein behaves as an unstructured and intrinsically disordered protein. In particular, α-syn carboxy-terminal (C-terminal) region exists in a disordered conformation in monomeric, fibrillar and membrane-bound states [23, 26–28] and plays an important role in the control of α-syn proprieties, notably its interactions with other proteins [29–34], metal ions [35, 36] and other ligands (e.g. dopamine and polyamines) [37].

Interestingly, α-syn C-terminal tail contains the majority of PTMs sites, including phosphorylation (Y125, S129, Y133 and Y136), truncation (D115, D119, P120, E130 and D135), ubiquitination (K96) and tissue transglutaminase cross-linking (Q109) [8, 38] (Fig. 1), suggesting that these PTMs may regulate α-syn structure and physiological functions. In this section, I will discuss how phosphorylation at S129 may act as a molecular switch for the regulation of α-syn functions and proteostasis.

Phosphorylation at S129 modulates α-syn-membrane binding

α-syn interaction with vesicles of different lipid compositions has been extensively studied (see reviews [39–41]) and recently, a great deal of attention has been focused on the role of phosphorylation at S129 in the regulation of α-syn-membrane interaction.

In vitro assessment of pS129 effect on α-syn-membrane binding has yielded controversial observations. Whereas some studies reported the absence of effect of S129 phosphorylation on α-syn-membrane interaction [42, 43], two independent groups showed that GRK-mediated authentic phosphorylation and S129→E (mimicking the phosphorylation state) decrease α-syn affinity to bind phospholipids [44, 45]. On the other hand, cellular and animal studies, based on the use of the phospho-mimic strategy, reported an inhibitory effect of S129 phosphorylation on α-syn-membrane interaction. In yeast and worm models of PD, S129→A substitution (to block phosphorylation) increases α-syn membrane-bound fraction, however the phospho-mimic mutation (S129D) inhibits its association with membranes [46, 47]. A similar observation has been reported in an adeno-associated virus (AAV)-based rat genetic model of PD where immuno-electron microscopy analysis detected the majority of the non-phosphorylatable α-syn mutant (S129A) associated with cellular membranes [48]. Collectively, in vitro and in vivo data support the hypothesis of an inhibitory effect of S129 phosphorylation on α-syn membrane binding.

It is worth noting that a non-negligible fraction of α-syn in cells is associated with membranes (i.e. mitochondrial membrane, synaptic vesicles) [39, 49] and this α-syn-associated fraction is also subjected to phosphorylation by membrane-associated kinases, namely the G protein-coupled receptor kinases (GRKs) [49]. A recently described role of membrane-associated α-syn phosphorylation is the regulation of neurotransmitter uptake, notably the dopamine [49]. In a cell-based assay, Hara and colleagues reported that GRK-mediated S129 phosphorylation enhances the ability of α-syn to increase dopamine uptake, without affecting cell surface expression of dopamine transporter [49]. These data suggest that pS129 plays an important role in the regulation of α-syn functions at the synaptic terminals. However, a question remains open on how phosphorylated α-syn modulates synaptic plasticity despite its low level [12] and short physiological life [43]. Hints for answering this question has been in part provided by Hirai and collaborators who reported that α-syn phosphorylation state is tightly regulated by physiological stimuli (i.e. stress) and proposed that pS129 may play an important role in the stress-induced synaptic plasticity [50].

Phosphorylation at S129 enhances α-syn interaction with metal ions

Several studies reported that α-syn C-terminal region is implicated in its interaction with metal ions [35, 36, 51] and the localization of S129 in a very close proximity to the putative metal binding sites [52–54] suggests that phosphorylation at this residue may modulate α-syn-metal ions binding (Fig. 1).

The effect of S129 phosphorylation on metal binding has been investigated in vitro using synthetic C-terminal peptide. Using Terbium (Tb3 +) as a luminescent probe of metal binding and isothermal titration calorimetry, Liu and colleagues showed that S129 phosphorylation has no effect on α-syn peptide (residues 119–132) interaction with trivalent ions [52]. In a more recent study, the use of a larger peptide fragment corresponding to the entire α-syn C-terminal region (residues 107–140) confirmed the absence of pS129 effect on α-syn interaction with trivalent ions, however it showed an increased binding affinities to divalent ions (Cu2 +, Pb2 + and Fe2 +) [53]. Moreover, tandem mass spectrometry analysis revealed that S129 phosphorylation affects ions binding sites. For example, the residue D119 involved in the binding of bivalent ions (Fe2 + and Pb2 +) to the non-phosphorylated peptide is not implicated in the interaction with pS129 peptide [53]. This observation suggests that pS129 may significantly affect α-syn conformation and redistribute metal ions binding sites.

It is widely confirmed that interaction with metal ions promotes α-syn fibrillization [55–57] and the question on how phosphorylation at S129 affects this process remains to be explored. In a recent study, Nubling and collaborators reported that mimicking phosphorylation by S129→E substitution facilitates oligomer formation in the presence of trivalent metal ions (Fe3 + and Al3 +), as compared to wild type protein [45]. This result is in contradiction with the observations described above reporting that pS129 has no effect on α-syn interaction with trivalent metal ions [52, 53]. This apparent discrepancy could be due to the fact that the authentically phosphorylated peptide and the phospho-mimic peptide may behave differently in the presence of metal ions. In summary, these results suggest that phosphorylation at S129 play an important role in the regulation of α-syn interaction with metal ions and could significantly affect metal ions-mediated α-syn structure and aggregation properties.

Phosphorylation at S129 regulates α-syn turnover

While several PTMs, notably ubiquitination [9, 58–61], sumoylation [62] and phosphorylation at Y39 [9, 63, 64], have been reported to regulate α-syn degradation via different proteolytic pathways, little is known about the implication of the phosphorylation at S129 on α-syn turnover. The first evidence of a cross-talk between phosphorylation at S129 and α-syn degradation has been reported by Chau and collaborators [65]. In their study, the authors observed that inhibition of the ubiquitin-proteasome system induced a significant increase of pS129 levels in human neuroblastoma [65]. A subsequent study confirmed this observation and reported that blocking the autophagy-lysosomal degradation pathway also induces a massive accumulation of pS129 in human neuroblastoma and rat cortical primary cultures [66]. Using a pulse-chase analysis and de novo protein synthesis inhibitor (cycloheximide), the authors observed that pS129 half-life time is significantly shorter (t½  = 54.9±6.4 min) compared to the non-phosphorylated form (t½ >240 min), suggesting that the phosphorylated form is selectively targeted for degradation [66].

More recently, our group reported that the overexpression of Polo-like kinase 2 (PLK2), the main kinase responsible for α-syn phosphorylation in the brain [67–69], enhances α-syn turnover via the autophagic degradation pathway [70]. This cell process is unique to the synuclein family (α and β-syn) and is governed by PLK2 kinase activity and by the direct interaction between PLK2 and α-syn [70] (Fig. 2). A similar observation has been reported in a yeast model of PD where S129→A substitution compromised the clearance of α-syn via the autophagic degradation pathway [71]. Although the physiological relevance of PLK2 and α-syn interaction remains unknown, a number of converging evidence suggest a synergistic role of these two proteins in the regulation of the synaptic transmission [6, 72–74] and cell response to oxidative stress [75–77]. Of note, our finding describing PLK2-mediated α-syn clearance offers new opportunities for the development of therapeutic strategies for the treatment of PD aiming at reducing, in a specific manner, the toxic levels of α-syn.

Phosphorylation at S129 modulates α-syn protein–protein interaction

The implication of α-syn C-terminal tail on its interaction with different proteins partners [29–34] raised the question on how phosphorylation at S129 may regulate α-syn interactome. McFarland and collaborators were the first to address this question using targeted functional proteomics approaches [78]. In this study, the authors showed that the non-phosphorylated α-syn peptide mainly interacts with proteins related to mitochondrial electron transport (complex I, III and IV proteins of the electron transport chain) [78], however the phosphorylated peptide had more affinity to certain cytoskeletal proteins and presynaptic proteins implicated in the synapse transmission and vesicle trafficking [78]. In a recent study, Yin and collaborators showed that the C-terminal region is implicated in α-syn interaction with Rab GTPases (Rab8a), a small guanine nucleotide binding proteins implicated in coordinating vesicle trafficking [79]. Using a battery of biophysical and cell culture approaches, the authors reported that phosphorylation at S129 promotes α-syn binding to Rab8a and modulates Rab8a-mediated α-syn toxicity [79]. These observations suggest that pS129 could serve as a molecular switch to control α-syn interaction with different protein partners and therefore modulates its functions. However, further investigations are required to assess the impact and the physiological consequences of S129 phosphorylation on α-syn interaction with other proteins, such as SNARE proteins [80, 81], cytoskeletal proteins (i.e. tubulin) [82, 83] and other amyloidogenic proteins (i.e. tau) [31, 32]

Phosphorylation at S129 regulates α-syn subcellular localization

α-syn exhibits different subcellular localizations (nuclear, cytoplasmic, neurites), suggesting that this protein may play a specific role in each cell compartment. This subcellular localization is regulated by different factors notably PTMs (i.e. monoubiquitination [84]), PD-linked mutations [85], protein sequence (i.e. α-syn N-terminal region) [85, 86] and implication of proteins partners (i.e. importin αa [86]). Despite the absence of nuclear targeting sequence, a growing number of cell-based assays and in vivo studies reported an important accumulation of phosphorylated α-syn in the nucleus. In cell culture, using anti-pS129 antibodies and immunocytochemistry approaches, several groups detected α-syn in the nucleus of mammalian cell lines [67, 87] and primary neuronal culture [67, 88, 89]. In vivo, α-syn nuclear localization has been observed in α-syn-transgenic mice [90–94], AAV-based rat model of PD [48] and α-syn-transgenic drosophila [16]. Furthermore, a biochemical approach has also helped in detecting pS129 expression in the nuclear fraction extracted from mice brains [95, 96]. Together, these observations suggest that phosphorylation at S129 may play a central role in the control of α-syn nuclear translocation.

This hypothesis was confirmed by Goncalves and Outeiro using a photoactivation-based approach to track α-syn intracellular dynamics in cell culture [85]. In their study, the authors observed a continuous α-syn trafficking between the cytoplasm and nucleus, which is in part regulated by α-syn phosphorylation state. Indeed, S129→A substitution to block phosphorylation significantly reduces α-syn nuclear translocation [85], suggesting that pS129 may act as a tag for α-syn targeting to the nuclear compartment.

Moreover, recent observations suggested that phosphorylation might act in synergistic pairs with PD-linked mutations to control α-syn nuclear localization. In a cell-based assay, Fares and collaborators showed that the newly described PD-linked mutation, G51D, is associated with an increase of pS129 levels and its accumulation in the nucleus of mammalian cells and primary neuronal cultures [97].

Interestingly, some specific kinases catalyzing α-syn phosphorylation at S129 may also play a role in pS129-mediated α-syn trafficking between the cytoplasm and nucleus. In cell culture, overexpression of GRK5 promotes α-syn translocation to the nucleus, while PLKs (PLK2 and PLK3) potentiate α-syn trafficking from the nucleus to the cytoplasm [67, 85].

It is important to note that studies of pS129 subcellular localization present certain limitations: 1) some anti-pS129 antibodies exhibit non-specific cross-reactivity to other unknown antigens present in the nucleus, since certain antibodies show a nuclear pS129 signal in α-syn-knockout tissue [96]; 2) the majority of α-syn subcellular localization analysis was performed in α-syn overexpressing systems, either transfected cells, viral based gene delivery or in transgenic animals. These observations raise the question on the physiological relevance of the used systems, and urge the development of optimized models to decorticate the exact role of S129 phosphorylation on α-syn nuclear translocation, its role in this subcellular compartment and its impact on cell survival.

Implication of α-syn phosphorylation at S129 in synucleinopathies pathogenesis and treatments

Abnormal accumulation of pS129 in synucleinopathy-diseased brain [12, 15] and the increase of its levels during ageing [99, 100], the greatest risk factor for PD, suggest that this PTM may represent a key player in the pathogenesis of PD and related disorders. However, the exact implication of S129 phosphorylation on α-syn aggregation and toxicity in vivo remains under debate.

Does phosphorylation enhance or suppress α-syn toxicity in vivo?

Investigation of the relative implication of pS129 on α-syn toxicity in vivo has yielded controversial results [8, 48, 101–103]. This apparent controversy is in part due to the fact that phospho-mimics (S129D/E) do not replicate the exact properties of the authentically phosphorylated α-syn [8, 42, 66]. Faced with this limitation, our group and others sought to address this question by overexpressing α-syn with its natural kinases, directly in the rat brain. Among the kinases responsible for the α-syn phosphorylation at S129 (casein kinases [14], the G protein-coupled receptor kinases (GRKs) [44], LRRK2 [104] and Polo-like kinases (PLKs) [67, 68]), two independent groups investigated the effect of the overexpression of GRK2 and GRK6 on α-syn toxicity in fly [18] and rodent genetic models of PD [105], respectively. In both models, GRKs overexpression was associated with an increase of pS129 levels and an enhanced cellular loss [18, 105]. It is important to note that GRK6-mediated phosphorylation moderately increased α-syn toxicity, however it significantly accelerated this process [105]. Together these observations suggest that GRK-mediated phosphorylation of α-syn exacerbates its toxicity in vivo. In a more recent study, our group reported an opposing result after overexpressing α-syn with another kinase, PLK2 [70]. In this study, we showed that AAV-mediated overexpression of PLK2 in the rat midbrain induced a 3-fold increase of pS129 levels in the infected neurons, a significant reduction of α-syn-associated dopaminergic neuronal loss and an alleviation of the hemi-parkinsonian motor impairment [70]. This effect is governed by PLK2 kinase activity and α-syn phosphorylation at S129 [70]. At the molecular level, our study demonstrated that PLK2 overexpression enhances α-syn clearance via autophagic degradation pathway, and suggest that PLK2-mediated neuroprotection effect is probably due to the reduction of intra-neuronal α-syn protein levels under the toxic threshold [70] (Fig. 2). The discrepancy between PLK2- and GRKs-mediated effects on α-syn toxicity could be in part due to the superiority of PLK2 to efficiently phosphorylate α-syn in vivo [70, 106]. Moreover, PLK2-mediated α-syn turnover depend on PLK2 and α-syn protein-protein interaction, suggesting that PLK2 may play the role of a co-chaperone to assist α-syn autophagic clearance. Collectively, these in vivo data demonstrate that the effect of phosphorylation on α-syn toxicity is governed by the kinase responsible for its phosphorylation, suggesting that kinases rather than the phosphorylation per se are key regulators of α-syn toxicity in vivo.

Is phosphorylation at S129 required for α-syn aggregation and seeding in vivo?

The impact of S129 phosphorylation on α-syn aggregation has been extensively studied in vitro and the data support the hypothesis of an inhibitory effect this PTM may have on α-syn fibrillogenesis [8, 42] (Fig. 2). However, the question on whether pS129 controls α-syn aggregation and seeding in vivo is still elusive.

In mammalian cell lines and primary neuronal culture, the addition of small amounts of exogenous α-syn pre-formed fibrils (Pffs) seeds the formation of intracellular α-syn aggregates recapitulating the main features of LBs [107, 108]. This cellular process is governed by protein seeding, a nucleation-dependent mechanism in which the α-syn Pffs provide a template for the assembly of soluble monomeric α-syn and lead to the formation of highly ordered protein aggregates [109]. Using this cell-based assay, Luk and colleagues showed that phosphorylation is not required for the formation of intracellular LB-like inclusions [107]. In this study, the authors transduced cells with α-syn S129A Pffs in cells stably overexpressing full length α-syn. Strikingly, immunocytochemistry revealed the presence of intracellular inclusions resembling those formed after wild type α-syn transduction, suggesting that phosphorylation at S129 is not required for inclusions seeding [107]. Moreover, the addition of truncated α-syn (1–120) Pffs to cells stably overexpressing wild type α-syn induced the formation of pS129-positive inclusions [107, 108]. Interestingly, this phospho-signal, which does not occur within the truncated Pffs, originates from the newly recruited endogenous α-syn and implies that phosphorylation may occur after inclusion formation [107, 108]. Collectively, these data demonstrate that phosphorylation is not a limiting factor for α-syn aggregation and seeding in vivo, raising the question on whether accumulation of pS129 is an early or a late event in synucleinopathies pathogenesis.

When does phosphorylation at S129 occur during PD pathogenesis: An early or late event?

Until today, the available data concerning pS129 accumulation in the brain have been collected in post-mortem tissues and the question on whether pS129 accumulation occurs during the early or late stages of synucleinopathies remains ambiguous.

In a recent work, Walker and collaborators investigated how pS129 levels and solubility change in cingulate and temporal cortex of DLB patients, at different stages of the disease. Using biochemical analysis, the authors reported a progressive accumulation of pS129-immunoractive species in diseased brains, compared to the healthy controls, and a positive correlation between pS129 levels and the severity of the disease symptoms [110]. Moreover, accumulation of insoluble phosphorylated forms, as well as the formation of pS129-postive insoluble species became detectable only at the late stages of the disease (stage IV and V, according to the Unified Staging System [111]) [110]. A similar study, using brain samples form patients suffering from PD, also reported a dramatic accumulation of pS129-positive inclusions in different brain regions at the late stages of the disease [112]. Together, these results demonstrate that abnormal accumulation of insoluble α-syn phosphorylated forms is mainly observed at the advanced stages of synucleinopathies, and suggest that accumulation of this PTM could be a late event in the disease progression. Moreover, the co-localization of α-syn with several kinases within LBs, notably CK2 [113], GRK5 [114] and LRRK2 [115] and the ability of these kinases to efficiently phosphorylate fibrillar and aggregated forms of α-syn [67, 116, 117], suggest that these kinases may also catalyze α-syn phosphorylation after LB formation.

Together these findings support the hypothesis that α-syn phosphorylation may occur after LBs formation and suggest that pS129 accumulation in the brain could represent a late event in the disease progression. It is plausible that the aberrant phosphorylation of α-syn within LBs may reflect an active process whereby phosphorylation may promote LBs disaggregation and/or enhance their clearance and degradation (Fig. 2)

Phosphorylation at S129 is a reliable biomarker for the diagnosis of PD and related disorders

In the past few years, the detection of phosphorylated α-syn at S129 in human cerebral spinal fluid (CSF) and blood plasma has been considered as a promising potential biomarker for the diagnosis of PD and related disorders [118]. Since then, several groups investigated the potential utility of pS129 accumulation in human fluids and peripheral nervous system as a biomarker for PD and synucleinopathies.

pS129 accumulation in human fluids and peripheral nervous system is the pathological hallmark of synucleinopathies

In 2013, Foulds and colleagues conducted a longitudinal study to investigate the relative changes of pS129 levels in the blood plasma of patients suffering from PD [119]. In this study, the authors showed that, although the levels of total α-syn were similar between PD patients and control subjects, pS129 levels were significantly higher in PD samples [119]. Moreover, statistical analysis confirmed the utility of pS129 plasma levels in discriminating patients with PD from healthy controls. Furthermore, recent studies reported that detection of phosphorylated α-syn inclusions within structures of the peripheral nervous system might also be a useful diagnostic test for PD and related synucleinopathies. Using skin biopsies, two independent groups reported pS129 accumulation in small and large nerve fibers in the majority of patients suffering from PD, while no signal was detected in healthy controls [120, 121]. Importantly, this cutaneous pathology was correlated with the evolution of the disease symptoms, suggesting that this peripheral marker can reflect the disease progression and could serve as a biomarker to monitor the disease evolution [120]. Moreover, other studies reported the presence of pS129-immunoreactive signal in gastric, duodenal and colonic biopsies [122, 123]. This pS129 pathology in gastric biopsies can be detected several years prior to the onset of motor symptoms, and therefore could be used for the early diagnosis of PD cases [123].

Collectively, these observations demonstrate that detection of pS129 in human fluids or in the peripheral nervous system could offer new opportunities for the development of promising biomarkers for the diagnosis of synucleinopathies and for monitoring their progression.

Accumulation of pS129 discriminates between synucleinopathy cases

Besides it potential role as a biomarker for PD and related disorders, converging evidence suggest that pS129 could also discriminate between the different synucleinopathy affections. In a post-mortem study, analysis of phosphorylated α-syn levels revealed a significant accumulation of pS129 in DLB-diseased brains, compared to healthy controls, confirming the association of pS129 accumulation and synucleinopathies [124]. Importantly, this study also revealed a difference in pS129 levels and solubility between the different synucleinopathies. In addition to its aberrant accumulation in PD and DLB-diseased brains, pS129 levels are higher in the PD with dementia (PDD) and DLB groups compared to PD without dementia (PDND), suggesting that pS129 accumulation is closely associated with demented cases [124].

Furthermore, pS129 levels in human fluids and in the peripheral nervous system could also discriminate between the different synucleinopathies cases. In a post-mortem study, Foulds and colleagues observed that the concentration of insoluble pS129 in cerebrospinal fluids is significantly higher in MSA samples compared to the other synucleinopathies, including PD and DLB [125]. More recently, pre-mortem analysis of pS129 signal in skin sympathetic nerve fibres and dermal nerve fibres revealed an accumulation of phosphorylated α-syn in PD patients, while no signal was detected in MSA or essential tremor control subjects. Together, these observations demonstrate that accumulation of pS129, rather then total α-syn, might provide a reliable test to discriminate between synucleinopathies and can help with the patients’ classification for a better management and recruitment for clinical trials.

Conclusions

Elucidation of the relative implication of S129 phosphorylation on α-syn aggregation, LBs formation and neurotoxicity is crucial for the understanding of synucleinopathies pathogenesis and the development of new disease-modifying treatments for PD and related disorders. However, some important questions remain unexplored: 1) identification of the exact α-syn physiological functions regulated by phosphorylation at S129. For example, a number of studies suggested the implication of pS129 in the control of α-syn subcellular localization [92, 93] and the modulation of dopamine synthesis [126], however the exact mechanism underlying these processes are not yet clear; 2) systematic investigation of the role of each kinase on α-syn neurotoxicity, with a focus on PLK2-mediated α-syn clearance as a viable target for the development of new therapeutic strategies for PD; 3) investigation of the cross talk between pS129 and other PTMs, for instance the interaction with Y125 phosphorylation. Indeed, recent studies suggested that S129 phosphorylation is tightly controlled by phosphorylation at Y125 [127] and it may regulate its toxicity in vivo [128]; 4) investigation of the synergistic role of PD-linked mutations and pS129 in the regulation of α-syn toxicity. This question was raised by a recent observation showing that E46K increases α-syn phosphorylation at S129 in cell culture, yeast and rodent models of PD [98]; 5) evaluation of the electrophysiological consequences of α-syn phosphorylation and its exact role in the regulation of stress-related synaptic plasticity; 6) assessment of the role of pS129 accumulation in the increased susceptibility of some brain regions to α-syn pathology. This question is motivated by the in vivo observation reporting a disparity in phosphorylated α-syn expression levels between different brain regions [50] and finally 7) investigation of the possible implications of phosphorylation on α-syn cell-to-cell transmission and its pathological propagation in PD-diseased brains. Ultimately, answers to these questions will lead to the identification of novel and more tractable therapeutic targets for the treatment of PD and related synucleinopathies.

CONFLICT OF INTEREST

The author has none to declare.

ACKNOWLEDGMENTS

This work was supported by the CHU de Quebec Foundation and CHU de Quebec Research Center-Université Laval. I would like to thank Ms. Razan Sheta for the critical review of this manuscript.

REFERENCES

[1] 

Lang AE , & Lozano AM ((1998) ) Parkinson’s disease. First of two parts. N Engl J Med, 339: , 1044–1053.

[2] 

Lang AE , & Lozano AM ((1998) ) Parkinson’s disease. Second of two parts. N Engl J Med, 339: , 1130–1143.

[3] 

McCann H , Stevens CH , Cartwright H , & Halliday GM ((2014) ) alpha-Synucleinopathy phenotypes. Parkinsonism Relat Disord, 20: (Suppl 1), S62–S67.

[4] 

Spillantini MG , Crowther RA , Jakes R , Cairns NJ , Lantos PL , & Goedert M ((1998) ) Filamentous alpha-synuclein inclusions link multiple system atrophy with Parkinson’s disease and dementia with Lewy bodies. Neurosci Lett, 251: , 205–208.

[5] 

Spillantini MG , Crowther RA , Jakes R , Hasegawa M , & Goedert M ((1998) ) alpha-Synuclein in filamentous inclusions of Lewy bodies from Parkinson’s disease and dementia with Lewy bodies. Proc Natl Acad Sci U S A, 95: , 6469–6473.

[6] 

Lashuel HA , Overk CR , Oueslati A , & Masliah E ((2013) ) The many faces of alpha-synuclein: From structure and toxicity to theraeutic target. Nat Rev Neurosci, 14: , 38–48.

[7] 

Lee VM , & Trojanowski JQ ((2006) ) Mechanisms of Parkinson’s disease linked to pathological alpha-synuclein: New targets for drug discovery. Neuron, 52: , 33–38.

[8] 

Oueslati A , Fournier M , & Lashuel HA ((2010) ) Role of post-translational modifications in modulating the structure, function and toxicity of alpha-synuclein: Imlications for Parkinson’s disease pathogenesis and therapies. Prog Brain Res, 183: , 115–145.

[9] 

Tenreiro S , Eckermann K , & Outeiro TF ((2014) ) Protein phosphorylation in neurodegeneration: Friend or foe? Front Mol Neurosci, 7: , 42.

[10] 

Muntane G , Ferrer I , & Martinez-Vicente M ((2012) ) alpha-synuclein phosphorylation and truncation are normal events in the adult human brain. Neuroscience, 200: , 106–119.

[11] 

Hasegawa M , Fujiwara H , Nonaka T , Wakabayashi K , Takahashi H , Lee VM , Trojanowski JQ , Mann D , & Iwatsubo T ((2002) ) Phosphorylated alpha-synuclein is ubiquitinated in alpha-synucleinopathy lesions. J Biol Chem, 277: , 49071–49076.

[12] 

Fujiwara H , Hasegawa M , Dohmae N , Kawashima A , Masliah E , Goldberg MS , Shen J , Takio K , & Iwatsubo T ((2002) ) alpha-Synuclein is phosphorylated in synucleinopathy lesions. Nat Cell Biol, 4: , 160–164.

[13] 

Kahle PJ , Neumann M , Ozmen L , & Haass C ((2000) ) Physiology and pathophysiology of alpha-synuclein. Cell culture and transgenic animal models based on a Parkinson’s disease-associated protein. Ann N Y Acad Sci, 920: , 33–41.

[14] 

Okochi M , Walter J , Koyama A , Nakajo S , Baba M , Iwatsubo T , Meijer L , Kahle PJ , & Haass C ((2000) ) Constitutive phosphorylation of the Parkinson’s disease associated alpha-synuclein. J Biol Chem, 275: , 390–397.

[15] 

Anderson JP , Walker DE , Goldstein JM , de Laat R , Banducci K , Caccavello RJ , Barbour R , Huang J , Kling K , Lee M , Diep L , Keim PS , Shen X , Chataway T , Schlossmacher MG , Seubert P , Schenk D , Sinha S , Gai WP , & Chilcote TJ ((2006) ) Phosphorylation of Ser-129 is the dominant pathological modification of alpha-synuclein in familial and sporadic Lewy body disease. J Biol Chem, 281: , 29739–29752.

[16] 

Takahashi M , Kanuka H , Fujiwara H , Koyama A , Hasegawa M , Miura M , & Iwatsubo T ((2003) ) Phosphorylation of alpha-synuclein characteristic of synucleinopathy lesions is recapitulated in alpha-synuclein transgenic Drosophila. Neurosci Lett, 336: , 155–158.

[17] 

Yamada M , Iwatsubo T , Mizuno Y , & Mochizuki H ((2004) ) Overexpression of alpha-synuclein in rat substantia nigra results in loss of dopaminergic neurons, phosphorylation of alpha-synuclein and activation of caspase-9: Resemblance to pathogenetic changes in Parkinson’s disease. J Neurochem, 91: , 451–461.

[18] 

Chen L , & Feany MB ((2005) ) Alpha-synuclein phosphorylation controls neurotoxicity and inclusion formation in a Drosophila model of Parkinson disease. Nat Neurosci, 8: , 657–663.

[19] 

Neumann M , Kahle PJ , Giasson BI , Ozmen L , Borroni E , Spooren W , Muller V , Odoy S , Fujiwara H , Hasegawa M , Iwatsubo T , Trojanowski JQ , Kretzschmar HA , & Haass C ((2002) ) Misfolded proteinase K-resistant hyperphosphorylated alpha-synuclein in aged transgenic mice with locomotor deterioration and in human alpha-synucleinopathies. J Clin Invest, 110: , 1429–1439.

[20] 

Bartels T , Choi JG , & Selkoe DJ ((2011) ) alpha-Synuclein occurs physiologically as a helically folded tetramer that resists aggregation. Nature, 477: , 107–110.

[21] 

Fauvet B , Fares MB , Samuel F , Dikiy I , Tandon A , Eliezer D , & Lashuel HA ((2012) ) Characterization of semisynthetic and naturally Nalpha-acetylated alpha-synuclein in vitro and in intact cells: Implications for aggregation and cellular properties of alpha-synuclein. J Biol Chem, 287: , 28243–28262.

[22] 

Dedmon MM , Lindorff-Larsen K , Christodoulou J , Vendruscolo M , & Dobson CM ((2005) ) Mapping long-range interactions in alpha-synuclein using spin-label NMR and ensemble molecular dynamics simulations. J Am Chem Soc, 127: , 476–477.

[23] 

Wu KP , Kim S , Fela DA , & Baum J ((2008) ) Characterization of conformational and dynamic properties of natively unfolded human and mouse alpha-synuclein ensembles by NMR: Implication for aggregation. J Mol Biol, 378: , 1104–1115.

[24] 

Burre J , Vivona S , Diao J , Sharma M , Brunger AT , & Sudhof TC ((2013) ) Properties of native brain alpha-synuclein. Nature, 498: , E4–6; discussion E6-7.

[25] 

Waudby CA , Camilloni C , Fitzpatrick AW , Cabrita LD , Dobson CM , Vendruscolo M , & Christodoulou J ((2013) ) In-cell NMR characterization of the secondary structure populations of a disordered conformation of alpha-synuclein within E. coli cells. PLoS One, 8: , e72286.

[26] 

Bertini I , Gupta YK , Luchinat C , Parigi G , Peana M , Sgheri L , & Yuan J ((2007) ) Paramagnetism-based NMR restraints provide maximum allowed probabilities for the different conformations of partially independent protein domains. J Am Chem Soc, 129: , 12786–12794.

[27] 

Eliezer D , Kutluay E , Bussell R Jr , & Browne G ((2001) ) Conformational properties of alpha-synuclein in its free and lipid-associated states. J Mol Biol, 307: , 1061–1073.

[28] 

Del Mar C , Greenbaum EA , Mayne L , Englander SW , & Woods VL Jr ((2005) ) Structure and properties of alpha-synuclein and other amyloids determined at the amino acid level. Proc Natl Acad Sci U S A, 102: , 15477–15482.

[29] 

Cherny D , Hoyer W , Subramaniam V , & Jovin TM ((2004) ) Double-stranded DNA stimulates the fibrillation of alpha-synuclein in vitro and is associated with the mature fibrils: An electron microscopy study. J Mol Biol, 344: , 929–938.

[30] 

Fernandez CO , Hoyer W , Zweckstetter M , Jares-Erijman EA , Subramaniam V , Griesinger C , & Jovin TM ((2004) ) NMR of alpha-synuclein-polyamine complexes elucidates the mechanism and kinetics of induced aggregation. EMBO J, 23: , 2039–2046.

[31] 

Giasson BI , Forman MS , Higuchi M , Golbe LI , Graves CL , Kotzbauer PT , Trojanowski JQ , & Lee VM ((2003) ) Initiation and synergistic fibrillization of tau and alpha-synuclein. Science, 300: , 636–640.

[32] 

Jensen PH , Hager H , Nielsen MS , Hojrup P , Gliemann J , & Jakes R ((1999) ) alpha-synuclein binds to Tau and stimulates the protein kinase A-catalyzed tau phosphorylation of serine residues 262 and 356. J Biol Chem, 274: , 25481–25489.

[33] 

Jensen PH , Islam K , Kenney J , Nielsen MS , Power J , & Gai WP ((2000) ) Microtubule-associated protein 1B is a component of cortical Lewy bodies and binds alpha-synuclein filaments. J Biol Chem, 275: , 21500–21507.

[34] 

Yap TL , Gruschus JM , Velayati A , Westbroek W , Goldin E , Moaven N , Sidransky E , & Lee JC ((2011) ) Alpha-synuclein interacts with Glucocerebrosidase providing a molecular link between Parkinson and Gaucher diseases. J Biol Chem, 286: , 28080–28088.

[35] 

Brown DR ((2007) ) Interactions between metals and alpha-synuclein–function or artefact? FEBS J, 274: , 3766–3774.

[36] 

Paik SR , Shin HJ , Lee JH , Chang CS , & Kim J ((1999) ) Copper(II)-induced self-oligomerization of alpha-synuclein. Biochem J, 340: (Pt 3), 821–828.

[37] 

Hoyer W , Cherny D , Subramaniam V , & Jovin TM ((2004) ) Impact of the acidic C-terminal region comprising amino acids 109-140 on alpha-synuclein aggregation in vitro. Biochemistry, 43: , 16233–16242.

[38] 

Schmid AW , Chiappe D , Pignat V , Grimminger V , Hang I , Moniatte M , & Lashuel HA ((2009) ) Dissecting the mechanisms of tissue transglutaminase-induced cross-linking of alpha-synuclein: Imlications for the pathogenesis of Parkinson disease. J Biol Chem, 284: , 13128–13142.

[39] 

Snead D , & Eliezer D ((2014) ) Alpha-synuclein function and dysfunction on cellular membranes. Exp Neurobiol, 23: , 292–313.

[40] 

Dikiy I , & Eliezer D ((2012) ) Folding and misfolding of alpha-synuclein on membranes. Biochim Biophys Acta, 1818: , 1013–1018.

[41] 

Pfefferkorn CM , Jiang Z , & Lee JC ((2012) ) Biophysics of alpha-synuclein membrane interactions. Biochim Biophys Acta, 1818: , 162–171.

[42] 

Paleologou KE , Schmid AW , Rospigliosi CC , Kim HY , Lamberto GR , Fredenburg RA , Lansbury PT Jr , Fernandez CO , Eliezer D , Zweckstetter M , & Lashuel HA ((2008) ) Phosphorylation at Ser-129 but not the phosphomimics S129E/D inhibits the fibrillation of alpha-synuclein. J Biol Chem, 283: , 16895–16905.

[43] 

Visanji NP , Wislet-Gendebien S , Oschipok LW , Zhang G , Aubert I , Fraser PE , & Tandon A ((2011) ) Effect of Ser-129 phosphorylation on interaction of alpha-synuclein with synaptic and cellular membranes. J Biol Chem, 286: , 35863–35873.

[44] 

Pronin AN , Morris AJ , Surguchov A , & Benovic JL ((2000) ) Synucleins are a novel class of substrates for G protein-coupled receptor kinases. J Biol Chem, 275: , 26515–26522.

[45] 

Nubling GS , Levin J , Bader B , Lorenzl S , Hillmer A , Hogen T , Kamp F , & Giese A ((2014) ) Modelling Ser129 phosphorylation inhibits membrane binding of pore-forming alpha-synuclein oligomers. PLoS One, 9: , e98906.

[46] 

Fiske M , Valtierra S , Solvang K , Zorniak M , White M , Herrera S , Konnikova A , Brezinsky R , & Debburman S ((2011) ) Contribution of Alanine-76 and Serine Phosphorylation in alpha-Synuclein Membrane Association and Aggregation in Yeasts. Parkinsons Dis, 2011: , 392180.

[47] 

Kuwahara T , Tonegawa R , Ito G , Mitani S , & Iwatsubo T ((2012) ) Phosphorylation of alpha-synuclein protein at Ser-129 reduces neuronal dysfunction by lowering its membrane binding property in Caenorhabditis elegans. J Biol Chem, 287: , 7098–7109.

[48] 

Azeredo da Silveira S , Schneider BL , Cifuentes-Diaz C , Sage D , Abbas-Terki T , Iwatsubo T , Unser M , & Aebischer P ((2009) ) Phosphorylation does not prompt, nor prevent, the formation of alpha-synuclein toxic species in a rat model of Parkinson’s disease. Hum Mol Genet, 18: , 872–887.

[49] 

Hara S , Arawaka S , Sato H , Machiya Y , Cui C , Sasaki A , Koyama S , & Kato T ((2013) ) Serine 129 phosphorylation of membrane-associated alpha-synuclein modulates dopamine transporter function in a G protein-coupled receptor kinase-dependent manner. Mol Biol Cell, 24: , 1649–1660, S1641-1643.

[50] 

Hirai Y , Fujita SC , Iwatsubo T , & Hasegawa M ((2004) ) Phosphorylated alpha-synuclein in normal mouse brain. FEBS Lett, 572: , 227–232.

[51] 

Bisaglia M , Tessari I , Mammi S , & Bubacco L ((2009) ) Interaction between alpha-synuclein and metal ions, still looking for a role in the pathogenesis of Parkinson’s disease. Neuromolecular Med, 11: , 239–251.

[52] 

Liu LL , & Franz KJ ((2007) ) Phosphorylation-dependent metal binding by alpha-synuclein peptide fragments. J Biol Inorg Chem, 12: , 234–247.

[53] 

Lu Y , Prudent M , Fauvet B , Lashuel HA , & Girault HH ((2011) ) Phosphorylation of alpha-Synuclein at Y125 and S129 alters its metal binding properties: Implications for understanding the role of alpha-Synuclein in the pathogenesis of Parkinson’s Disease and related disorders. ACS Chem Neurosci, 2: , 667–675.

[54] 

Binolfi A , Rasia RM , Bertoncini CW , Ceolin M , Zweckstetter M , Griesinger C , Jovin TM , & Fernandez CO ((2006) ) Interaction of alpha-synuclein with divalent metal ions reveals key differences: A link between structure, binding specificity and fibrillation enhancement. J Am Chem Soc, 128: , 9893–9901.

[55] 

Kostka M , Hogen T , Danzer KM , Levin J , Habeck M , Wirth A , Wagner R , Glabe CG , Finger S , Heinzelmann U , Garidel P , Duan W , Ross CA , Kretzschmar H , & Giese A ((2008) ) Single particle characterization of iron-induced pore-forming alpha-synuclein oligomers. J Biol Chem, 283: , 10992–11003.

[56] 

Hogen T , Levin J , Schmidt F , Caruana M , Vassallo N , Kretzschmar H , Botzel K , Kamp F , & Giese A ((2012) ) Two different binding modes of alpha-synuclein to lipid vesicles depending on its aggregation state. Biophys J, 102: , 1646–1655.

[57] 

Levin J , Hogen T , Hillmer AS , Bader B , Schmidt F , Kamp F , Kretzschmar HA , Botzel K , & Giese A ((2011) ) Generation of ferric iron links oxidative stress to alpha-synuclein oligomer formation. J Parkinsons Dis, 1: , 205–216.

[58] 

Rott R , Szargel R , Haskin J , Bandopadhyay R , Lees AJ , Shani V , & Engelender S ((2011) ) alpha-Synuclein fate is determined by USP9X-regulated monoubiquitination. Proc Natl Acad Sci U S A, 108: , 18666–18671.

[59] 

Haj-Yahya M , Fauvet B , Herman-Bachinsky Y , Hejjaoui M , Bavikar SN , Karthikeyan SV , Ciechanover A , Lashuel HA , & Brik A ((2013) ) Synthetic polyubiquitinated alpha-Synuclein reveals important insights into the roles of the ubiquitin chain in regulating its pathophysiology. Proc Natl Acad Sci U S A, 110: , 17726–17731.

[60] 

Tofaris GK , Kim HT , Hourez R , Jung JW , Kim KP , & Goldberg AL ((2011) ) Ubiquitin ligase Nedd4 promotes alpha-synuclein degradation by the endosomal-lysosomal pathway. Proc Natl Acad Sci U S A, 108: , 17004–17009.

[61] 

Lonskaya I , Desforges NM , Hebron ML , & Moussa CE ((2013) ) Ubiquitination increases parkin activity to promote autophagic alpha-synuclein clearance. PLoS One, 8: , e83914.

[62] 

Shahpasandzadeh H , Popova B , Kleinknecht A , Fraser PE , Outeiro TF , & Braus GH ((2014) ) Interplay between sumoylation and phosphorylation for protection against alpha-synuclein inclusions. J Biol Chem, 289: , 31224–31240.

[63] 

Mahul-Mellier AL , Fauvet B , Gysbers A , Dikiy I , Oueslati A , Georgeon S , Lamontanara AJ , Bisquertt A , Eliezer D , Masliah E , Halliday G , Hantschel O , & Lashuel HA ((2014) ) c-Abl phosphorylates alpha-synuclein and regulates its degradation: Implication for alpha-synuclein clearance and contribution to the pathogenesis of Parkinson’s disease. Hum Mol Genet, 23: , 2858–2879.

[64] 

Hebron ML , Lonskaya I , & Moussa CE ((2013) ) Nilotinib reverses loss of dopamine neurons and improves motor behavior via autophagic degradation of alpha-synuclein in Parkinson’s disease models. Hum Mol Genet, 22: , 3315–3328.

[65] 

Chau KY , Ching HL , Schapira AH , & Cooper JM ((2009) ) Relationship between alpha synuclein phosphorylation, proteasomal inhibition and cell death: Relevance to Parkinson’s disease pathogenesis. J Neurochem, 110: , 1005–1013.

[66] 

Machiya Y , Hara S , Arawaka S , Fukushima S , Sato H , Sakamoto M , Koyama S , & Kato T ((2010) ) Phosphorylated alpha-synuclein at Ser-129 is targeted to the proteasome pathway in a ubiquitin-independent manner. J Biol Chem, 285: , 40732–40744.

[67] 

Mbefo MK , Paleologou KE , Boucharaba A , Oueslati A , Schell H , Fournier M , Olschewski D , Yin G , Zweckstetter M , Masliah E , Kahle PJ , Hirling H , & Lashuel HA ((2010) ) Phosphorylation of synucleins by members of the Polo-like kinase family. J Biol Chem, 285: , 2807–2822.

[68] 

Inglis KJ , Chereau D , Brigham EF , Chiou SS , Schobel S , Frigon NL , Yu M , Caccavello RJ , Nelson S , Motter R , Wright S , Chian D , Santiago P , Soriano F , Ramos C , Powell K , Goldstein JM , Babcock M , Yednock T , Bard F , Basi GS , Sham H , Chilcote TJ , McConlogue L , Griswold-Prenner I , & Anderson JP ((2009) ) Polo-like kinase 2 (PLK2) phosphorylates alpha-synuclein at serine 129 in central nervous system. J Biol Chem, 284: , 2598–2602.

[69] 

Bergeron M , Motter R , Tanaka P , Fauss D , Babcock M , Chiou SS , Nelson S , San Pablo F , & Anderson JP ((2014) ) In vivo modulation of polo-like kinases supports a key role for PLK2 in Ser129 alpha-synuclein phosphorylation in mouse brain. Neuroscience, 256: , 72–82.

[70] 

Oueslati A , Schneider BL , Aebischer P , & Lashuel HA ((2013) ) Polo-like kinase 2 regulates selective autophagic alpha-synuclein clearance and suppresses its toxicity in vivo. Proc Natl Acad Sci U S A, 110: , E3945–E3954.

[71] 

Tenreiro S , Reimao-Pinto MM , Antas P , Rino J , Wawrzycka D , Macedo D , Rosado-Ramos R , Amen T , Waiss M , Magalhaes F , Gomes A , Santos CN , Kaganovich D , & Outeiro TF ((2014) ) Phosphorylation modulates clearance of alpha-synuclein inclusions in a yeast model of Parkinson’s disease. PLoS Genet, 10: , e1004302.

[72] 

Seeburg DP , Pak D , & Sheng M ((2005) ) Polo-like kinases in the nervous system. Oncogene, 24: , 292–298.

[73] 

Seeburg DP , Feliu-Mojer M , Gaiottino J , Pak DT , & Sheng M ((2008) ) Critical role of CDK5 and Polo-like kinase 2 in homeostatic synaptic plasticity during elevated activity. Neuron, 58: , 571–583.

[74] 

Burre J ((2015) ) The Synaptic Function of alpha-Synuclein. J Parkinsons Dis, 5: (4, 699–713.

[75] 

Matsumoto T , Wang PY , Ma W , Sung HJ , Matoba S , & Hwang PM ((2009) ) Polo-like kinases mediate cell survival in mitochondrial dysfunction. Proc Natl Acad Sci U S A, 106: , 14542–14546.

[76] 

Li J , Ma W , Wang PY , Hurley PJ , Bunz F , & Hwang PM ((2014) ) Polo-like kinase 2 activates an antioxidant pathway to promote the survival of cells with mitochondrial dysfunction. Free Radic Biol Med, 73: , 270–277.

[77] 

Hashimoto M , Hsu LJ , Rockenstein E , Takenouchi T , Mallory M , & Masliah E ((2002) ) alpha-Synuclein protects against oxidative stress via inactivation of the c-Jun N-terminal kinase stress-signaling pathway in neuronal cells. J Biol Chem, 277: , 11465–11472.

[78] 

McFarland MA , Ellis CE , Markey SP , & Nussbaum RL ((2008) ) Proteomics analysis identifies phosphorylation-dependent alpha-synuclein protein interactions. Mol Cell Proteomics, 7: , 2123–2137.

[79] 

Yin G , Lopes da Fonseca T , Eisbach SE , Anduaga AM , Breda C , Orcellet ML , Szego EM , Guerreiro P , Lazaro DF , Braus GH , Fernandez CO , Griesinger C , Becker S , Goody RS , Itzen A , Giorgini F , Outeiro TF , & Zweckstetter M ((2014) ) alpha-Synuclein interacts with the switch region of Rab8a in a Ser129 phosphorylation-dependent manner. Neurobiol Dis, 70: , 149–161.

[80] 

Burre J , Sharma M , Tsetsenis T , Buchman V , Etherton MR , & Sudhof TC ((2010) ) Alpha-synuclein promotes SNARE-complex assembly in vivo and in vitro. Science, 329: , 1663–1667.

[81] 

Burre J , Sharma M , & Sudhof TC ((2012) ) Systematic mutagenesis of alpha-synuclein reveals distinct sequence requirements for physiological and pathological activities. J Neurosci, 32: , 15227–15242.

[82] 

Payton JE , Perrin RJ , Clayton DF , & George JM ((2001) ) Protein-protein interactions of alpha-synuclein in brain homogenates and transfected cells. Brain Res Mol Brain Res, 95: , 138–145.

[83] 

Zhou RM , Huang YX , Li XL , Chen C , Shi Q , Wang GR , Tian C , Wang ZY , Jing YY , Gao C , & Dong XP ((2010) ) Molecular interaction of alpha-synuclein with tubulin influences on the polymerization of microtubule in vitro and structure of microtubule in cells. Mol Biol Rep, 37: , 3183–3192.

[84] 

Monti B , Polazzi E , Batti L , Crochemore C , Virgili M , & Contestabile A ((2007) ) Alpha-synuclein protects cerebellar granule neurons against 6-hydroxydopamine-induced death. J Neurochem, 103: , 518–530.

[85] 

Goncalves S , & Outeiro TF ((2013) ) Assessing the subcellular dynamics of alpha-synuclein using photoactivation microscopy. Mol Neurobiol, 47: , 1081–1092.

[86] 

Ma KL , Song LK , Yuan YH , Zhang Y , Han N , Gao K , & Chen NH ((2014) ) The nuclear accumulation of alpha-synuclein is mediated by importin alpha and promotes neurotoxicity by accelerating the cell cycle. Neuropharmacology, 82: , 132–142.

[87] 

Lee BR , Matsuo Y , Cashikar AG , & Kamitani T ((2013) ) Role of Ser129 phosphorylation of alpha-synuclein in melanoma cells. J Cell Sci, 126: , 696–704.

[88] 

McLean PJ , Ribich S , & Hyman BT ((2000) ) Subcellular localization of alpha-synuclein in primary neuronal cultures: Effect of missense mutations. J Neural Transm Suppl, 53–63.

[89] 

Seo JH , Rah JC , Choi SH , Shin JK , Min K , Kim HS , Park CH , Kim S , Kim EM , Lee SH , Lee S , Suh SW , & Suh YH ((2002) ) Alpha-synuclein regulates neuronal survival via Bcl-2 family expression and PI3/Akt kinase pathway. FASEB J, 16: , 1826–1828.

[90] 

Masliah E , Rockenstein E , Veinbergs I , Mallory M , Hashimoto M , Takeda A , Sagara Y , Sisk A , & Mucke L ((2000) ) Dopaminergic loss and inclusion body formation in alpha-synuclein mice: Implications for neurodegenerative disorders. Science, 287: , 1265–1269.

[91] 

Wakamatsu M , Ishii A , Ukai Y , Sakagami J , Iwata S , Ono M , Matsumoto K , Nakamura A , Tada N , Kobayashi K , Iwatsubo T , & Yoshimoto M ((2007) ) Accumulation of phosphorylated alpha-synuclein in dopaminergic neurons of transgenic mice that express human alpha-synuclein. J Neurosci Res, 85: , 1819–1825.

[92] 

Amschl D , Neddens J , Havas D , Flunkert S , Rabl R , Romer H , Rockenstein E , Masliah E , Windisch M , & Hutter-Paier B ((2013) ) Time course and progression of wild type alpha-synuclein accumulation in a transgenic mouse model. BMC Neurosci, 14: , 6.

[93] 

Schell H , Hasegawa T , Neumann M , & Kahle PJ ((2009) ) Nuclear and neuritic distribution of serine-129 phosphorylated alpha-synuclein in transgenic mice. Neuroscience, 160: , 796–804.

[94] 

Goers J , Manning-Bog AB , McCormack AL , Millett IS , Doniach S , Di Monte DA , Uversky VN , & Fink AL ((2003) ) Nuclear localization of alpha-synuclein and its interaction with histones. Biochemistry, 42: , 8465–8471.

[95] 

Yu S , Li X , Liu G , Han J , Zhang C , Li Y , Xu S , Liu C , Gao Y , Yang H , Ueda K , & Chan P ((2007) ) Extensive nuclear localization of alpha-synuclein in normal rat brain neurons revealed by a novel monoclonal antibody. Neuroscience, 145: , 539–555.

[96] 

Huang Z , Xu Z , Wu Y , & Zhou Y ((2011) ) Determining nuclear localization of alpha-synuclein in mouse brains. Neuroscience, 199: , 318–332.

[97] 

Fares MB , Ait-Bouziad N , Dikiy I , Mbefo MK , Jovicic A , Kiely A , Holton JL , Lee SJ , Gitler AD , Eliezer D , & Lashuel HA ((2014) ) The novel Parkinson’s disease linked mutation G51D attenuates in vitro aggregation and membrane binding of alpha-synuclein, and enhances its secretion and nuclear localization in cells. Hum Mol Genet, 23: , 4491–4509.

[98] 

Mbefo MK , Fares MB , Paleologou K , Oueslati A , Yin G , Tenreiro S , Pinto M , Outeiro T , Zweckstetter M , Masliah E , & Lashuel HA ((2015) ) Parkinson disease mutant E46K enhances alpha-synuclein phosphorylation in mammalian cell lines, in yeast, and in vivo. J Biol Chem, 290: , 9412–9427.

[99] 

McCormack AL , Mak SK , & Di Monte DA ((2012) ) Increased alpha-synuclein phosphorylation and nitration in the aging primate substantia nigra. Cell Death Dis, 3: , e315.

[100] 

Canron MH , Perret M , Vital A , Bezard E , & Dehay B ((2012) ) Age-dependent alpha-synuclein aggregation in the Microcebus murinus lemur primate. Sci Rep, 2: , 910.

[101] 

McFarland NR , Fan Z , Xu K , Schwarzschild MA , Feany MB , Hyman BT , & McLean PJ ((2009) ) Alpha-synuclein S129 phosphorylation mutants do not alter nigrostriatal toxicity in a rat model of Parkinson disease. J Neuropathol Exp Neurol, 68: , 515–524.

[102] 

Gorbatyuk OS , Li S , Sullivan LF , Chen W , Kondrikova G , Manfredsson FP , Mandel RJ , & Muzyczka N ((2008) ) The phosphorylation state of Ser-129 in human alpha-synuclein determines neurodegeneration in a rat model of Parkinson disease. Proc Natl Acad Sci U S A, 105: , 763–768.

[103] 

Febbraro F , Sahin G , Farran A , Soares S , Jensen PH , Kirik D , & Romero-Ramos M ((2013) ) Ser129D mutant alpha-synuclein induces earlier motor dysfunction while S129A results in distinctive pathology in a rat model of Parkinson’s disease. Neurobiol Dis, 56: , 47–58.

[104] 

Qing H , Wong W , McGeer EG , & McGeer PL ((2009) ) Lrrk2 phosphorylates alpha synuclein at serine 129: Parkinson disease implications. Biochem Biophys Res Commun, 387: , 149–152.

[105] 

Sato H , Arawaka S , Hara S , Fukushima S , Koga K , Koyama S , & Kato T ((2011) ) Authentically phosphorylated alpha-synuclein at Ser129 accelerates neurodegeneration in a rat model of familial Parkinson’s disease. J Neurosci, 31: , 16884–16894.

[106] 

Salvi M , Trashi E , Marin O , Negro A , Sarno S , & Pinna LA ((2012) ) Superiority of PLK-2 as alpha-synuclein phosphorylating agent relies on unique specificity determinants. Biochem Biophys Res Commun, 418: , 156–160.

[107] 

Luk KC , Song C , O’Brien P , Stieber A , Branch JR , Brunden KR , Trojanowski JQ , & Lee VM ((2009) ) Exogenous alpha-synuclein fibrils seed the formation of Lewy body-like intracellular inclusions in cultured cells. Proc Natl Acad Sci U S A, 106: , 20051–20056.

[108] 

Volpicelli-Daley LA , Luk KC , Patel TP , Tanik SA , Riddle DM , Stieber A , Meaney DF , Trojanowski JQ , & Lee VM ((2011) ) Exogenous alpha-synuclein fibrils induce Lewy body pathology leading to synaptic dysfunction and neuron death. Neuron, 72: , 57–71.

[109] 

Oueslati A , Ximerakis M , & Vekrellis K ((2014) ) Protein Transmission, Seeding and Degradation: Key Steps for alpha-Synuclein Prion-Like Propagation. Exp Neurobiol, 23: , 324–336.

[110] 

Walker DG , Lue LF , Adler CH , Shill HA , Caviness JN , Sabbagh MN , Akiyama H , Serrano GE , Sue LI , Beach TG , & Arizona Parkinson Disease C ((2013) ) Changes in properties of serine 129 phosphorylated alpha-synuclein with progression of Lewy-type histopathology in human brains. Exp Neurol, 240: , 190–204.

[111] 

Beach TG , Adler CH , Lue L , Sue LI , Bachalakuri J , Henry-Watson J , Sasse J , Boyer S , Shirohi S , Brooks R , Eschbacher J , White CL 3rd , Akiyama H , Caviness J , Shill HA , Connor DJ , Sabbagh MN , Walker DG , & Arizona Parkinson’s Disease C ((2009) ) Unified staging system for Lewy body disorders: Correlation with nigrostriatal degeneration, cognitive impairment and motor dysfunction. Acta Neuropathol, 117: , 613–634.

[112] 

Zhou J , Broe M , Huang Y , Anderson JP , Gai WP , Milward EA , Porritt M , Howells D , Hughes AJ , Wang X , & Halliday GM ((2011) ) Changes in the solubility and phosphorylation of alpha-synuclein over the course of Parkinson’s disease. Acta Neuropathol, 121: , 695–704.

[113] 

Ryu MY , Kim DW , Arima K , Mouradian MM , Kim SU , & Lee G ((2008) ) Localization of CKII beta subunits in Lewy bodies of Parkinson’s disease. J Neurol Sci, 266: , 9–12.

[114] 

Arawaka S , Wada M , Goto S , Karube H , Sakamoto M , Ren CH , Koyama S , Nagasawa H , Kimura H , Kawanami T , Kurita K , Tajima K , Daimon M , Baba M , Kido T , Saino S , Goto K , Asao H , Kitanaka C , Takashita E , Hongo S , Nakamura T , Kayama T , Suzuki Y , Kobayashi K , Katagiri T , Kurokawa K , Kurimura M , Toyoshima I , Niizato K , Tsuchiya K , Iwatsubo T , Muramatsu M , Matsumine H , & Kato T ((2006) ) The role of G-protein-coupled receptor kinase 5 in pathogenesis of sporadic Parkinson’s disease. J Neurosci, 26: , 9227–9238.

[115] 

Qing H , Zhang Y , Deng Y , McGeer EG , & McGeer PL ((2009) ) Lrrk2 interaction with alpha-synuclein in diffuse Lewy body disease. Biochem Biophys Res Commun, 390: , 1229–1234.

[116] 

Waxman EA , & Giasson BI ((2011) ) Characterization of kinases involved in the phosphorylation of aggregated alpha-synuclein. J Neurosci Res, 89: , 231–247.

[117] 

Oueslati A , Paleologou KE , Schneider BL , Aebischer P , & Lashuel HA ((2012) ) Mimicking phosphorylation at serine 87 inhibits the aggregation of human alpha-synuclein and protects against its toxicity in a rat model of Parkinson’s disease. J Neurosci, 32: , 1536–1544.

[118] 

Foulds PG , Mitchell JD , Parker A , Turner R , Green G , Diggle P , Hasegawa M , Taylor M , Mann D , & Allsop D ((2011) ) Phosphorylated alpha-synuclein can be detected in blood plasma and is potentially a useful biomarker for Parkinson’s disease. FASEB J, 25: , 4127–4137.

[119] 

Foulds PG , Diggle P , Mitchell JD , Parker A , Hasegawa M , Masuda-Suzukake M , Mann DM , & Allsop D ((2013) ) A longitudinal study on alpha-synuclein in blood plasma as a biomarker for Parkinson’s disease. Sci Rep, 3: , 2540.

[120] 

Doppler K , Ebert S , Uceyler N , Trenkwalder C , Ebentheuer J , Volkmann J , & Sommer C ((2014) ) Cutaneous neuropathy in Parkinson’s disease: A window into brain pathology. Acta Neuropathol, 128: , 99–109.

[121] 

Donadio V , Incensi A , Leta V , Giannoccaro MP , Scaglione C , Martinelli P , Capellari S , Avoni P , Baruzzi A , & Liguori R ((2014) ) Skin nerve alpha-synuclein deposits: A biomarker for idiopathic Parkinson disease. Neurology, 82: , 1362–1369.

[122] 

Pouclet H , Lebouvier T , Coron E , Neunlist M , & Derkinderen P ((2012) ) Lewy pathology in gastric and duodenal biopsies in Parkinson’s Disease. Mov Disor, 27: , 708.

[123] 

Hilton D , Stephens M , Kirk L , Edwards P , Potter R , Zajicek J , Broughton E , Hagan H , & Carroll C ((2014) ) Accumulation of alpha-synuclein in the bowel of patients in the pre-clinical phase of Parkinson’s disease. Acta Neuropathol, 127: , 235–241.

[124] 

Swirski M , Miners JS , de Silva R , Lashley T , Ling H , Holton J , Revesz T , & Love S ((2014) ) Evaluating the relationship between amyloid-beta and alpha-synuclein phosphorylated at Ser129 in dementia with Lewy bodies and Parkinson’s disease. Alzheimers Res Ther, 6: , 77.

[125] 

Foulds PG , Yokota O , Thurston A , Davidson Y , Ahmed Z , Holton J , Thompson JC , Akiyama H , Arai T , Hasegawa M , Gerhard A , Allsop D , & Mann DM ((2012) ) Post mortem cerebrospinal fluid alpha-synuclein levels are raised in multiple system atrophy and distinguish this from the other alpha-synucleinopathies, Parkinson’s disease and Dementia with Lewy bodies. Neurobiol Dis, 45: , 188–195.

[126] 

Lou H , Montoya SE , Alerte TN , Wang J , Wu J , Peng X , Hong CS , Friedrich EE , Mader SA , Pedersen CJ , Marcus BS , McCormack AL , Di Monte DA , Daubner SC , & Perez RG ((2010) ) Serine 129 phosphorylation reduces the ability of alpha-synuclein to regulate tyrosine hydroxylase and protein phosphatase 2A in vitro and in vivo. J Biol Chem, 285: , 17648–17661.

[127] 

Kosten J , Binolfi A , Stuiver M , Verzini S , Theillet FX , Bekei B , van Rossum M , & Selenko P ((2014) ) Efficient modification of alpha-synuclein serine 129 by protein kinase CK1 requires phosphorylation of tyrosine 125 as a priming event. ACS Chem Neurosci, 5: , 1203–1208.

[128] 

Chen L , Periquet M , Wang X , Negro A , McLean PJ , Hyman BT , & Feany MB ((2009) ) Tyrosine and serine phosphorylation of alpha-synuclein have opposing effects on neurotoxicity and soluble oligomer formation. J Clin Invest, 119: , 3257–3265.

[129] 

Ulmer TS , Bax A , Cole NB , & Nussbaum RL ((2005) ) Structure and dynamics of micelle-bound human alpha-synuclein. J Biol Chem, 280: , 9595–9603.

Figures and Tables

Fig.1

α-syn C-terminal residues subjected to post-translational modifications or implicated in α-syn interaction with metalions. Schematic representation of α-syn (Protein Data Bank ID: 1XQ8) [129] with the N-terminal region, the NAC region and the C-terminal region are colored in blue, orange and red, respectively. In the upper part of the scheme are represented the binding regions (amino acids in bold) with bivalent metal ions. In the lower part of the scheme are represented the potential sites, in α-syn C-terminal region, subjected to post-translational modifications, including phosphorylation, ubiquitination, tissue transglutaminase cross-linking and truncation.

α-syn C-terminal residues subjected to post-translational modifications or implicated in α-syn interaction with metalions. Schematic representation of α-syn (Protein Data Bank ID: 1XQ8) [129] with the N-terminal region, the NAC region and the C-terminal region are colored in blue, orange and red, respectively. In the upper part of the scheme are represented the binding regions (amino acids in bold) with bivalent metal ions. In the lower part of the scheme are represented the potential sites, in α-syn C-terminal region, subjected to post-translational modifications, including phosphorylation, ubiquitination, tissue transglutaminase cross-linking and truncation.
Fig.2

Potential role of phosphorylation at S129 in the regulation of α-syn clearance, aggregation and toxicity. Converging lines of evidence supports the implication of pS129 in the regulation of α-syn turnover. In our recent report we showed that PLK2 phosphorylates, interacts with and enhances α-syn clearance via the lysosome-autophagy degradation pathway [70], thereby it suppresses its toxicity in vivo. In another hand, in vitro assays revealed that authentic phosphorylation at S129, inhibits α-syn fibrillogenesis [42], suggesting that this post-translational modification may reduce α-syn fibrillogenesis and aggregation-related toxicity. Finally, α-syn oligomers and fibrils are good substrates for several kinases, notably PLK2 [116], suggesting that this event could occur after LBs formation. It is plausible that S129 phosphorylation may represent an active process promoting LBs disaggregation and/or clearance. Collectively, these observations suggest that phosphorylation at S129 may play a protective role against α-syn toxicity. PLKs: Polo like kinases, CK2: Casein kinase 2, GRKs: G protein-coupled receptor kinases, KD: Kinase domain, PBD: Polo box domain.

Potential role of phosphorylation at S129 in the regulation of α-syn clearance, aggregation and toxicity. Converging lines of evidence supports the implication of pS129 in the regulation of α-syn turnover. In our recent report we showed that PLK2 phosphorylates, interacts with and enhances α-syn clearance via the lysosome-autophagy degradation pathway [70], thereby it suppresses its toxicity in vivo. In another hand, in vitro assays revealed that authentic phosphorylation at S129, inhibits α-syn fibrillogenesis [42], suggesting that this post-translational modification may reduce α-syn fibrillogenesis and aggregation-related toxicity. Finally, α-syn oligomers and fibrils are good substrates for several kinases, notably PLK2 [116], suggesting that this event could occur after LBs formation. It is plausible that S129 phosphorylation may represent an active process promoting LBs disaggregation and/or clearance. Collectively, these observations suggest that phosphorylation at S129 may play a protective role against α-syn toxicity. PLKs: Polo like kinases, CK2: Casein kinase 2, GRKs: G protein-coupled receptor kinases, KD: Kinase domain, PBD: Polo box domain.