Next Article in Journal
Photosynthesis: Genetic Strategies Adopted to Gain Higher Efficiency
Previous Article in Journal
Role of miR-9 in Modulating NF-κB Signaling and Cytokine Expression in COVID-19 Patients
Previous Article in Special Issue
A-Syn(ful) MAM: A Fresh Perspective on a Converging Domain in Parkinson’s Disease
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Role of Lipids in the Pathogenesis of Parkinson’s Disease

1
Department of Neurology, Kindai University Faculty of Medicine, Osaka-Sayama 589-8511, Osaka, Japan
2
Department of Neurology, Osaka University Graduate School of Medicine, Suita 565-0871, Osaka, Japan
3
Life Science Research Institute, Kindai University, Osaka-Sayama 589-8511, Osaka, Japan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(16), 8935; https://doi.org/10.3390/ijms25168935
Submission received: 8 July 2024 / Revised: 7 August 2024 / Accepted: 10 August 2024 / Published: 16 August 2024
(This article belongs to the Special Issue The Structure and Function of Synuclein)

Abstract

:
Aggregation of α-synuclein (αSyn) and its accumulation as Lewy bodies play a central role in the pathogenesis of Parkinson’s disease (PD). However, the mechanism by which αSyn aggregates in the brain remains unclear. Biochemical studies have demonstrated that αSyn interacts with lipids, and these interactions affect the aggregation process of αSyn. Furthermore, genetic studies have identified mutations in lipid metabolism-associated genes such as glucocerebrosidase 1 (GBA1) and synaptojanin 1 (SYNJ1) in sporadic and familial forms of PD, respectively. In this review, we focus on the role of lipids in triggering αSyn aggregation in the pathogenesis of PD and propose the possibility of modulating the interaction of lipids with αSyn as a potential therapy for PD.

1. Introduction

Parkinson’s disease (PD) is a neurodegenerative disease in which patients show motor symptoms, such as tremor, bradykinesia, and rigidity, as well as nonmotor symptoms, such as dementia, autonomic dysfunction, and sleep disorders. Dopamine replacement therapy is used as a symptomatic treatment, but there is no treatment available at present that can attenuate the progression of PD [1]. Pathologically, PD is characterized by Lewy bodies (LBs) composed mainly of α-synuclein (αSyn), which are also characteristic features of dementia with Lewy bodies (DLB) [2,3,4]. Genetic studies have shown that missense mutations (A30P, E46K, H50Q, G51D, A53T, and A53E), as well as duplications and triplication mutations in the SNCA gene encoding αSyn, are responsible for familial forms of PD (fPD) [5,6,7,8,9,10,11,12,13]. Genome-wide association studies have also reported that single nucleotide polymorphisms (SNPs) in the SNCA gene are a major risk factor for sporadic PD (sPD) [14,15,16]. Moreover, animal studies have demonstrated that the expression of either wild-type (wt) or mutant αSyn induces neurodegeneration accompanied by the accumulation of αSyn aggregates, as well as motor function deficits [17,18,19]. These lines of evidence strongly suggest that αSyn plays a central role in the pathogenesis of both familial and sporadic forms of PD. Biochemical studies have shown that αSyn aggregates into oligomers and amyloid-like fibrils in vitro under various conditions [20,21]. Moreover, these αSyn oligomers and amyloid-like fibrils have been shown to exert toxicity. However, it is still unknown as to how wt αSyn aggregates in vivo despite its important roles in the pathogenesis of PD.
αSyn is expressed in the brain, which is a lipid-rich tissue [22,23]. Lipids in the brain include fatty acids, triacylglycerols, phospholipids, sterols, and glycolipids. Physiologically, αSyn is suggested to be involved in fatty acid metabolism in the brain [24,25]. Furthermore, αSyn is reported to interact with the lipid membrane. Interestingly, its interaction with lipids is shown to induce the aggregation of αSyn, implying the possible roles of lipids in the pathogenesis of PD [26]. Supporting this hypothesis, mutations in the glucocerebrosidase 1 (GBA1) and synaptojanin 1 (SYNJ1) genes, which are involved in lipid metabolism, have been identified as risk factors for sPD and fPD, respectively. We and others have demonstrated that dysregulation of the substrate lipids by GBA1 and SYNJ1 mutations promotes the aggregation of wt αSyn [27,28,29]. Furthermore, additional genetic risk factors, including phospholipase A2G6 (PLA2G6), vacuolar protein sorting 13C (VPS13C), chromosome 19 open reading frame 12 (C19orf12), and galactosylceramidase (GALC), have also been reported [30,31,32,33]. Moreover, using Fourier-transform infrared spectroscopy imaging, we demonstrated the accumulation of lipids in the central region of LBs in the brains of PD patients [34]. In this review, we focus on the association between lipids and αSyn aggregation. Furthermore, we propose that lipids play an important role in the pathogenesis of PD via the aggregation of αSyn.

2. Association between Lipids and αSyn Aggregation

Synucleins are a family that comprises the α, β, and γSyns. While αSyn is associated with diseases such as PD and DLB, no disease-associated mutations have been reported for βSyn and γSyn [35]. αSyn is expressed in several tissues, including the brain, erythrocytes, lymphocytes, muscle, kidney, heart, and lung. In the brain, αSyn has been reported to be localized in the presynaptic terminals of neurons [36]. αSyn binds to lipid membranes, particularly to synaptic vesicles in presynaptic terminals [36,37,38,39,40,41,42], and in the rat brain, about 15% of αSyn is in the membrane-bound form [43]. It was also reported that αSyn and βSyn colocalize in the mouse brain and human brain [41,42,44].
αSyn consists of 140 amino acids and is divided into the following three regions: the N-terminal region (amino acids 1–60), the non-amyloid-β component (NAC) region (amino acids 61–95), and the C-terminal region (amino acids 96–140) (Figure 1). αSyn is thought to exist in an unfolded state in aqueous solution [38,45]. On the other hand, αSyn is known to adopt an α-helix structure when bound to lipid membranes in the cellular milieu. Moreover, αSyn is converted to β-sheet-rich aggregates in LBs [34,46,47,48,49,50]. The NAC region is essential for the aggregation of αSyn, which is ultimately converted to a β-sheet-rich structure during the aggregation process [50,51,52]. Deletion of the highly hydrophobic amino acids 71 to 82 in the NAC domain of αSyn abolishes its aggregation [50], and peptides derived from the NAC region alone can form β-sheet-rich fibrils [53], indicating that the NAC domain of αSyn is important for aggregation. The N-terminal region of αSyn, unlike the C-terminal region, is highly conserved among the synuclein species [38,39,40]. The C-terminal region is known to interact with the NAC region, resulting in the suppression of αSyn aggregation [54]. Moreover, the C-terminal region of αSyn also transiently interacts with its N-terminal region to form a compact monomeric state, leading to the suppression of aggregation [55].
The physiological function of αSyn remains poorly understood. αSyn is abundant in presynaptic terminals and is known to colocalize with presynaptic proteins [56]. In addition, the overexpression of αSyn in mice impairs synaptic function [57,58,59,60]. All the mutations in αSyn that cause familial PD are located in the N-terminal lipid membrane-binding domain. Among them, overexpression of the A53T and E46K mutants of αSyn has been reported to inhibit neurotransmitter release [58,61,62]. The A53T mutant has preserved membrane-binding affinity, whereas the E46K mutant has enhanced membrane-binding affinity [61]. On the other hand, the overexpression of A30P, which has reduced membrane-binding affinity, does not show defects during exocytosis [41]. Taken together, these findings suggest that the membrane-binding capacity of the N-terminal region of αSyn is important for synaptic function. The C-terminal region of αSyn has been shown to bind directly to synaptobrevin-2 (VAMP2), a vesicle-associated soluble NSF-attachment protein receptor (v-SNARE), and αSyn is known to bind to the synaptic plasma membrane via VAMP2, promoting vesicle fusion and clustering [63,64]. In addition, αSyn has been reported to promote the expansion of fusion pores during exocytosis [65]. In addition, other studies have suggested that the binding of αSyn oligomers to the N-terminal region of VAMP2 inhibits toxic αSyn aggregation on synaptic vesicles [66]. These data indicate that αSyn may take different forms and function differently under different conditions.
The binding of αSyn to lipid membranes has been investigated using artificial liposomes in vitro [67]. Negatively charged phospholipids, such as phosphatidylserine (PS) and phosphatidic acid (PA), are known to affect this binding, as are lipids such as sphingolipids and fatty acids [61,68,69,70,71,72,73]. Phospholipids are components of biological membranes but are also important in cellular functions [23]. Previous studies have reported that the neutral phospholipids phosphatidylcholine and phosphatidylethanolamine do not interact with αSyn alone but with vesicles containing acidic phospholipids, such as PA and PS [67,69,72,74]. Interestingly, the lipids found to interact with αSyn in these biochemical experiments resemble synaptic vesicles in vivo and are abundant in PA and PS [75,76]. In addition to the head group, the polyunsaturation of acyl chains is also important for the binding of αSyn to the membrane, which influences the fluidity of the plasma membrane, and short saturated acyl chains have been reported to affect αSyn aggregation [68,72,77]. Moreover, previous studies have reported the interaction between lipid rafts and αSyn. The localization of αSyn at the synapse depends on its interaction with lipid rafts. Experiments using mouse brains have showed that the changes in the composition of lipid rafts and the binding affinity of αSyn may impair the localization of αSyn and the normal function of αSyn at the synapse [41]. The process by which wt αSyn, which is initially nontoxic, aggregates and gains toxicity is unclear at present, but lipids may be involved in this process; one hypothesis is that lipids induce a conformational change in αSyn that leads to the aggregation of αSyn [43]. The other hypothesis is that the lipid membrane promotes an increase in the local concentration of αSyn, which triggers its aggregation. Although αSyn is natively unfolded, upon binding to lipid membranes, the N-terminal region of αSyn undergoes a conformational change from a random coil to an α-helix structure [78,79,80]. However, it is yet unclear as to how membrane-bound αSyn gains a β-sheet structure and initiates aggregation.
In the pathogenesis of PD and DLB, αSyn is thought to change to a β-sheet-rich structure during the process of aggregation, which eventually results in LB formation [81]. Several studies have suggested the prion-like propagation of αSyn aggregates, in which αSyn takes on a pathological structure [82,83,84,85,86,87,88,89,90]. However, it remains unknown as to how αSyn, which is normally considered to be nontoxic, gains such a pathological structure. We hypothesize that the interaction of αSyn with lipid membranes may be one of the key steps in this process. One reason is that all the point mutations linked to familial PD are located in the N-terminal membrane-binding domain of αSyn. However, in vitro studies have shown an inconsistency in the binding affinity of these mutants to lipid membranes. For example, the A30P, A53E, and G51D mutants were reported to have a decreased affinity for lipid membranes compared with wild-type αSyn, the A53T and H50Q mutants had an unchanged affinity, and the E46K mutant had an increased affinity [41,61,69,74,91,92,93,94,95,96,97]. In addition, the A53T and E46K mutants were found to form fibrils with different structures in the absence of membranes [95], suggesting that lipid membranes affect the aggregation process of αSyn. The mechanisms of the conflicting effects of different αSyn mutations on lipid binding and aggregation of αSyn have not been fully elucidated (Table 1). Thus, further studies are needed to understand the relationship between the effects of different αSyn mutations on its lipid binding and aggregation. The second reason is that αSyn has been reported to change its structure from a random coil to an α-helix-containing conformation upon binding to lipid membranes, although this conformational change in αSyn was less efficient in the A30P mutant, which has a reduced affinity for lipid membranes [69]. Third, we demonstrated using Fourier-transform infrared microspectroscopy that lipids accumulate in the central region of LBs in the brains of PD patients [34]. Recently, using high-resolution microscopy, LBs in the brains of PD patients were shown to contain membrane fragments, vesicles, and organelles such as mitochondria and lysosomes [98]. Taken together, it is likely that the interaction between αSyn and lipid membranes is involved in the process of αSyn aggregation and LB formation.
It is also known that the binding of αSyn to membranes can adversely affect membrane integrity and cause deleterious effects. αSyn aggregates, particularly annular oligomers, induce membrane permeabilization like other amyloidogenic proteins such as amyloid-β (Aβ) and tau [99]. The protofibrillar form of αSyn, but not the monomeric and fibrillar forms, was shown to permeabilize synthetic vesicles in vitro [100]. The A30P and A53T mutants of αSyn, but not the G51D mutant, showed higher permeabilizing activities than wt αSyn, suggesting that the effects of mutation on membrane affinity do not always correlate with their permeabilizing activities [101,102].
The N-terminal region of αSyn is amphiphilic and is involved in the interaction with lipid membranes [36,103]. The KTKEGV repeat motif in the N-terminal region of αSyn is well conserved among species [104]. Interestingly, similar amphiphilic [DE]-[DE]-X-R-X-R-L-G repeat motifs are also found in apolipoproteins involved in lipid metabolism [36,105]. In fact, αSyn has similar characteristics to apolipoproteins in that it has amphipathic helices, is inserted into membranes, and affects membrane curvature [106], implying that αSyn may be a potential member of the apolipoprotein family [107]. Several apolipoproteins have been implicated in neurodegenerative diseases [108,109,110]. For example, previous studies have shown that increased levels of apolipoprotein E (ApoE) and its receptor LRP-1 are risk factors for PD. In particular, ApoEε4, which binds to lipids with a higher affinity than other ApoE isoforms, is the most pathogenic and promotes the aggregation of αSyn [111,112]. Paslawski et al. showed that αSyn and apolipoproteins colocalize on lipoprotein vesicles in cerebrospinal fluid, and that ApoE levels were increased in the cerebrospinal fluid of patients with early PD [113]. Considering the similarity between αSyn and apolipoproteins, αSyn may also be involved in the pathogenesis of PD via lipid interactions. As described above, missense mutations in αSyn that cause fPD are all located within and between the KTKEGV repeat motifs in the N-terminal region. These mutations have been shown to alter the interaction between αSyn and lipid membranes, resulting in the aggregation of αSyn.
Recently, protein liquid–liquid phase separation (LLPS) has received much attention as the potential trigger of pathological protein aggregation observed in neurodegenerative diseases. LLPS is a phenomenon in which a protein solution is separated into different liquid phases without mixing, resulting in the formation of liquid droplets. In cells, proteins as well as RNAs are concentrated by LLPS, which contributes to the formation of membraneless organelles, such as nucleoli, stress granules, and P-bodies [114,115,116,117,118,119,120]. Recently, αSyn, as well as other neurodegenerative disease-causing proteins, including FUS, TDP-43, and tau, have been shown to reversibly form liquid droplets by LLPS and to eventually form irreversible amyloid-like fibrils [121,122,123,124,125], suggesting that the liquid droplets formed by LLPS may serve as a precursor of pathological protein aggregation. αSyn is known to have two low-complexity domains (LCDs), which are prone to undergo LLPS, in the N-terminal and NAC regions. Previous studies have shown that both regions are important for the LLPS of αSyn. Moreover, Ray et al. have shown that the formation of liquid droplets of αSyn leads to the formation of its amyloid-like fibrils under various conditions [126]. The formation of liquid droplets and amyloid-like fibrils of αSyn is accelerated by pH changes, high concentrations, phosphate modifications, fPD mutations, metal ions (Fe3+, Cu2+, and Ca2+), and liposomes [126,127]. These results suggest that lipids could affect not only the LLPS of αSyn but also its aggregation. Interestingly, it has been suggested that the central region of LBs contains a large amount of lipids such as sphingomyelin [34,128]. Considering these findings, lipids may play an important role in triggering the aggregation of αSyn in the pathogenesis of PD (Figure 2).

3. αSyn and Lipids in Brain Tissue

The classical LBs found in the substantia nigra of PD patients are typically spherical intracellular inclusions that are eosinophilic. They have a dense core in the center, surrounded by a halo of radial fibrils that are approximately 10 nm in width [4,129]. On the other hand, cortical LBs found in the cortex of DLB and advanced PD patients are somewhat distinct and have no halo, unlike classical LBs [4,130]. The main component of LBs is αSyn. Other than αSyn, they are composed of neurofilaments [131], microtubule-associated protein 1B [132], and the galectin-3 [133] proteins, as well as various lipids [34].
In addition to the proteinaceous constituents, there have been reports on the presence of lipids in the LBs of PD patients. Histochemical lipid staining demonstrated the presence of phospholipids, particularly sphingomyelin, in the core of LBs [134]. Another study also reported positive staining with the lipid-soluble fluorescent dye rhodamine B in the core of LBs of PD patients [135]. Using Fourier-transform infrared microscopy, we found an accumulation of lipids that was surrounded by a halo rich in β-sheet fibrils in the core region of LBs in the brains of PD patients [34]. These observations described above are consistent with those of a recent study that used advanced correlative light and electron microscopy to demonstrate that LBs contain αSyn fibrils as well as membranous materials, such as various vesicles, mitochondria, and lysosomes [96].

4. Mutations in Lipid Metabolism-Associated Genes and PD Pathogenesis

The involvement of lipids in αSyn aggregation in the pathogenesis of PD is not only supported by the abovementioned lines of biochemical evidence but also by lines of genetic evidence. The GBA1 gene, which encodes the lysosomal enzyme glucocerebrosidase (GCase) and whose homozygous mutations cause Gaucher disease, is recognized as the strongest genetic risk factor for sPD [136,137]. GCase hydrolyzes glucosylceramide (GlcCer), which accumulates in Gaucher disease. To elucidate the mechanism by which mutations in the GBA1 gene increase the risk of developing PD, we conducted a combination of in vitro and in vivo analyses using a Drosophila model of PD expressing αSyn [138]. We found that knockdown of GBA1 accelerates the degeneration of dopamine neurons, resulting in motor dysfunction in these PD flies [28]. Furthermore, we found that the GBA1 substrate GlcCer directly acts on αSyn and accelerates the accumulation of proteinase K-resistant αSyn in vitro, suggesting that GlcCer promotes the toxic structural conversion of αSyn [28]. In addition, GlcCer has been shown to colocalize with αSyn in iPS-derived neurons [139]. Moreover, pathological studies have suggested a broad association between αSyn pathology and other lysosomal storage diseases in which other glycosphingolipids accumulate in brains [140,141], implying that such glycosphingolipids may also trigger αSyn aggregation.
Interestingly, decreased activity of the lysosomal enzyme GCase has been reported in the brains of sporadic PD/DLB patients [142,143]. In addition, lysosomal enzymes show high activity in the substantia nigra, and decreases in these activities have been noted in PD/DLB patients and in older individuals [144,145]. Taken together, these findings suggest that lysosomal enzyme activity may be involved in the etiology of PD/DLB.
Mutations in the SYNJ1 gene which encodes the phosphoinositide phosphatase SYNJ1 have been found to be associated with early-onset fPD (PARK20) [146]. Recently, we investigated the role of SYNJ1 in the pathogenesis of PD in relation to αSyn aggregation and found that SYNJ1 deficiency causes the accumulation of phosphatidylinositol-3,4,5-trisphosphate (PIP3), which directly interacts with αSyn and accelerates the formation of αSyn fibrils. Interestingly, these αSyn amyloid-like fibrils are similar in shape to the αSyn amyloid-like fibrils found in PD brains. Furthermore, using a C. elegans model of PD, SYNJ1 haploinsufficiency was found to accelerate αSyn accumulation and to induce locomotor defects. Moreover, we confirmed that PIP3 and αSyn colocalize in sPD brains [29]. These results suggest that the interaction of PIP3 with αSyn may contribute to the aggregation of αSyn in the pathogenesis of PD.
Moreover, other genetic variants, including PLA2G6 (PARK14), VPS13C, C19orf12, and GALC, also suggest an association between lipid abnormalities and PD [30,31,32,33]. The PLA2G6 gene is responsible for an autosomal recessive form of PD (PARK14), as well as infantile neuroaxonal dystrophy and neurodegeneration with iron accumulation in the brain, all of which show LB pathology [30]. PLA2G6 belongs to the phospholipase PLA2 family of proteins, which hydrolyze phospholipids and generate free fatty acids and lysophospholipids. PLA2G6 knockout mice were reported to show increased levels and the accumulation of αSyn in their neurons [147]. The loss of iPLA2-VIA, a homolog of PLA2G6, in Drosophila resulted in increased ceramide levels and resulted in a shortened lifespan and impaired synaptic transmission due to neurodegeneration [148]. Knockdown of Vps13, which is related to a lipid transport protein, in Drosophila was shown to increase αSyn oligomerization [149]. Mutations in the C19orf12 gene, which is involved in fatty acid supply, has also been reported to be associated with PD [32]. Krabbe’s disease, an inherited lysosomal storage disorder (LSD) caused by mutations in the lysosomal enzyme GALC, is also characterized by neuronal αSyn aggregates [150]. An experimental study demonstrated that a mouse model of Krabbe’s disease exhibits αSyn aggregation in the brain, altered lipid membrane dynamics, and impaired synaptic function and macroautophagy [33].
Gaucher disease is one of the LSDs in which the activity of specific lysosomal enzymes is defective, resulting in the accumulation of their substrates, such as specific lipids, glycoproteins, and mucopolysaccharides, within lysosomes. Interestingly, αSyn accumulation is found not only in the brains of GD patients but also in the brains of several other LSD patients [151,152,153]. For example, in β-galactosialidosis, GM1 ganglioside (GM1) accumulates in the brain, and GM1 was found to specifically bind to αSyn, promoting its aggregation [153]. We also showed that the knockdown of β-galactosidase promotes the formation of proteinase K-resistant αSyn aggregates and worsens locomotor dysfunction in Drosophila [28]. These results suggest that accumulated glycolipids interact with αSyn to promote its toxic conversion and subsequent neurodegeneration. It is noteworthy that GM1 also binds directly to the Aβ peptide and induces its toxic conversion [140]. In light of these findings, GM1 may be involved in the toxic conversion of disease-associated proteins to pathological structures that are common among these disease-associated proteins. Ganglioside levels were reported to be increased in the neurons of LSD patients, and the LB-positive LSD patients showed an accumulation of GlcCer, the gangliosides GM1, GM2, GM3, and SM, and cholesterol [141]. Biochemical studies showed the direct binding of αSyn to these lipids, and the binding affinities were as follows: GM3 > Gb3 > GalCer > GM1 > sulfatide > LacCer > GM4 > GM2 > asialo-GM1 > GD3, suggesting that glycolipids with 1, 3, 5 sugar units are preferred [154]. Grey et al. reported that GM1 and GM3 promote αSyn aggregation, and these glycolipids are rich within exosomes, implying that the nucleation of αSyn aggregation is triggered within exosomes [155]. Moreover, saturated fatty acids (SFA), monounsaturated fatty acids (MUFA), and polyunsaturated fatty acids (PUFA) are suggested to be an important factor in αSyn aggregation [156,157,158,159,160,161,162,163,164,165,166]. In addition, the modifications to lipids could also affect the αSyn aggregation process. For example, it has been suggested that lipid peroxidation is involved in the oligomerization of αSyn [167]. These multiple factors may be differentially involved in the impairment of the normal function αSyn. Taken together, altered lipid metabolism may cause the accumulation of lipids in the brain, and these accumulated lipids may promote αSyn aggregation, leading to neurodegeneration.

5. Conclusions

In this review, we focused on lipids as one of the key factors in triggering αSyn aggregation. Biochemical studies demonstrated the interaction of αSyn with lipids and the effects of these lipids on αSyn aggregation, and genetic studies also support this concept. Pathological studies also indicated the involvement of lipids in LBs in the brains of PD patients. Thus, we proposed here that the interaction of αSyn with lipids would be one of the key steps that triggers and/or induces αSyn aggregation, although the detailed mechanisms by which lipids promote αSyn aggregation remain to be elucidated. Therefore, we propose that modulating the interaction of lipids with αSyn using small molecules or by modulating the contents of lipids in food, for example, would be a new therapeutic option towards developing potential therapies for PD.

Author Contributions

Conceptualization, D.O. and Y.N.; writing—original draft preparation, S.K., D.O. and Y.N.; writing—review and editing, K.I. and Y.N.; supervision, Y.N.; funding acquisition, K.I. and Y.N. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported in part by Grants-in-Aid for Scientific Research (A) (24H00630 to Y.N.), for Scientific Research (B) (21H02840 to Y.N.), for Challenging Exploratory Research (17K19658 to Y.N.), and for Transformative Research Areas (A) (Multifaceted Proteins) (20H05927 to Y.N) from the Ministry of Education, Culture, Sports, Science, and Technology (MEXT), Japan; by grants for Practical Research Projects for Rare/Intractable Diseases (JP16ek0109018 and JP19ek0109222 to Y.N.) from the Japan Agency for Medical Research and Development; by Intramural Research Grants for Neurological and Psychiatric Disorders (30-3, 30-9, 3-9, and 6-9 to Y.N.) from the National Center of Neurology and Psychiatry; and by a grant from the Japan Foundation for Neuroscience and Mental Health, Japan (to Y.N.).

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

We thank H. Akiko Popiel for English proofreading and editing of the manuscript.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Bloem, B.R.; Okun, M.S.; Klein, C. Parkinson’s disease. Lancet 2021, 397, 2284–2303. [Google Scholar] [CrossRef] [PubMed]
  2. Spillantini, M.G.; Schmidt, M.L.; Lee, V.M.; Trojanowski, J.Q.; Jakes, R.; Goedert, M. α-Synuclein in Lewy bodies. Nature 1997, 388, 839–840. [Google Scholar] [CrossRef] [PubMed]
  3. Spillantini, M.G.; Crowther, R.A.; Jakes, R.; Cairns, N.J.; Lantos, P.L.; Goedert, M. Filamentous α-synuclein inclusions link multiple system atrophy with Parkinson’s disease and dementia with Lewy bodies. Neurosci. Lett. 1998, 251, 205–208. [Google Scholar] [CrossRef] [PubMed]
  4. 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] [PubMed]
  5. Krüger, R.; Kuhn, W.; Muller, T.; Woitalla, D.; Graeber, M.; Kosel, S.; Przuntek, H.; Epplen, J.T.; Schols, L.; Riess, O. Ala30Pro mutation in the gene encoding α-synuclein in Parkinson’s disease. Nat. Genet. 1998, 18, 106–108. [Google Scholar] [CrossRef] [PubMed]
  6. Zarranz, J.J.; Alegre, J.; Gomez-Esteban, J.C.; Lezcano, E.; Ros, R.; Ampuero, I.; Vidal, L.; Hoenicka, J.; Rodriguez, O.; Atares, B.; et al. The new mutation, E46K, of α-synuclein causes Parkinson and Lewy body dementia. Ann. Neurol. 2004, 55, 164–173. [Google Scholar] [CrossRef] [PubMed]
  7. Appel-Cresswell, S.; Vilarino-Guell, C.; Encarnacion, M.; Sherman, H.; Yu, I.; Shah, B.; Weir, D.; Thompson, C.; Szu-Tu, C.; Trinh, J.; et al. Alpha-synuclein p.H50Q, a novel pathogenic mutation for Parkinson’s disease. Mov. Disord. 2013, 28, 811–813. [Google Scholar] [CrossRef] [PubMed]
  8. Proukakis, C.; Dudzik, C.G.; Brier, T.; MacKay, D.S.; Cooper, J.M.; Millhauser, G.L.; Houlden, H.; Schapira, A.H. A novel α-synuclein missense mutation in Parkinson disease. Neurology 2013, 80, 1062–1064. [Google Scholar] [CrossRef] [PubMed]
  9. Lesage, S.; Anheim, M.; Letournel, F.; Bousset, L.; Honore, A.; Rozas, N.; Pieri, L.; Madiona, K.; Durr, A.; Melki, R.; et al. G51D α-synuclein mutation causes a novel parkinsonian-pyramidal syndrome. Ann. Neurol. 2013, 73, 459–471. [Google Scholar] [CrossRef]
  10. Polymeropoulos, M.H. Mutation in the α-synuclein gene identified in families with Parkinson’s disease. Science 1997, 276, 2045–2047. [Google Scholar] [CrossRef]
  11. Pasanen, P.; Myllykangas, L.; Siitonen, M.; Raunio, A.; Kaakkola, S.; Lyytinen, J.; Tienari, P.J.; Poyhonen, M.; Paetau, A. Novel α-synuclein mutation A53E associated with atypical multiple system atrophy and Parkinson’s disease-type pathology. Neurobiol. Aging 2014, 35, 2180.e1–2180.e5. [Google Scholar] [CrossRef]
  12. Singleton, A.B.; Farrer, M.; Johnson, J.; Singleton, A.; Hague, S.; Kachergus, J.; Hulihan, M.; Peuralinna, T.; Dutra, A.; Nussbaum, R.; et al. α-Synuclein locus triplication causes Parkinson’s disease. Science 2003, 302, 841. [Google Scholar] [CrossRef]
  13. Chartier-Harlin, M.C.; Kachergus, J.; Roumier, C.; Mouroux, V.; Douay, X.; Lincoln, S.; Levecque, C.; Larvor, L.; Andrieux, J.; Hulihan, M.; et al. α-Synuclein locus duplication as a cause of familial Parkinson’s disease. Lancet 2004, 364, 1167–1169. [Google Scholar] [CrossRef]
  14. Mueller, J.C.; Fuchs, J.; Hofer, A.; Zimprich, A.; Lichtner, P.; Illig, T.; Berg, D.; Wullner, U.; Meitinger, T.; Gasser, T. Multiple regions of α-synuclein are associated with Parkinson’s disease. Ann. Neurol. 2005, 57, 535–541. [Google Scholar] [CrossRef]
  15. Satake, W.; Nakabayashi, Y.; Mizuta, I.; Hirota, Y.; Ito, C.; Kubo, M.; Kawaguchi, T.; Tsunoda, T.; Watanabe, M.; Takeda, A.; et al. Genome-wide association study identifies common variants at four loci as genetic risk factors for Parkinson’s disease. Nat. Genet. 2009, 41, 1303–1307. [Google Scholar] [CrossRef]
  16. Simon-Sanchez, J.; Schulte, C.; Bras, J.M.; Sharma, M.; Gibbs, J.R.; Berg, D.; Paisan-Ruiz, C.; Lichtner, P.; Scholz, S.W.; Hernandez, D.G.; et al. Genome-wide association study reveals genetic risk underlying Parkinson’s disease. Nat. Genet. 2009, 41, 1308–1312. [Google Scholar] [CrossRef] [PubMed]
  17. Feany, M.B.; Bender, W.W. A Drosophila model of Parkinson’s disease. Nature 2000, 404, 394–398. [Google Scholar] [CrossRef]
  18. Lee, M.K.; Stirling, W.; Xu, Y.; Xu, X.; Qui, D.; Mandir, A.S.; Dawson, T.M.; Copeland, N.G.; Jenkins, N.A.; Price, D.L. Human α-synuclein-harboring familial Parkinson’s disease-linked Ala-53 → Thr mutation causes neurodegenerative disease with α-synuclein aggregation in transgenic mice. Proc. Natl. Acad. Sci. USA 2002, 99, 8968–8973. [Google Scholar] [CrossRef] [PubMed]
  19. Masliah, E.; Rockenstein, E.; Veinbergs, I.; Mallory, M.; Hashimoto, M.; Takeda, A.; Sagara, Y.; Sisk, A.; Mucke, L. Dopaminergic loss and inclusion body formation in α-synuclein mice: Implications for neurodegenerative disorders. Science 2000, 287, 1265–1269. [Google Scholar] [CrossRef] [PubMed]
  20. Conway, K.A.; Harper, J.D.; Lansbury, P.T. Accelerated in vitro fibril formation by a mutant α-synuclein linked to early-onset Parkinson disease. Nat. Med. 1998, 4, 1318–1320. [Google Scholar] [CrossRef]
  21. Uversky, V.N. Neuropathology, biochemistry, and biophysics of α-synuclein aggregation. J. Neurochem. 2007, 103, 17–37. [Google Scholar] [CrossRef] [PubMed]
  22. Sastry, P.S. Lipids of nervous tissue: Composition and metabolism. Prog. Lipid Res. 1985, 24, 69–176. [Google Scholar] [CrossRef] [PubMed]
  23. Farooqui, A.A.; Horrocks, L.A.; Farooqui, T. Glycerophospholipids in brain: Their metabolism, incorporation into membranes, functions, and involvement in neurological disorders. Chem. Phys. Lipids 2000, 106, 1–29. [Google Scholar] [CrossRef] [PubMed]
  24. Sorrentino, Z.A.; Giasson, B.I.; Chakrabarty, P. α-Synuclein and astrocytes: Tracing the pathways from homeostasis to neurodegeneration in Lewy body disease. Acta Neuropathol. 2019, 138, 1–21. [Google Scholar] [CrossRef] [PubMed]
  25. Castagnet, P.I.; Golovko, M.Y.; Barceló-Coblijn, G.C.; Nussbaum, R.L.; Murphy, E.J. Fatty acid incorporation is decreased in astrocytes cultured from alpha-synuclein gene-ablated mice. J Neurochem. 2005, 94, 839–849. [Google Scholar] [CrossRef] [PubMed]
  26. Galvagnion, C. The role of lipids interacting with α-synuclein in the pathogenesis of Parkinson’s disease. J. Park. Dis. 2017, 7, 433–450. [Google Scholar] [CrossRef] [PubMed]
  27. Sidransky, E.; Lopez, G. The link between the GBA gene and parkinsonism. Lancet Neurol. 2012, 11, 986–998. [Google Scholar] [CrossRef] [PubMed]
  28. Suzuki, M.; Fujikake, N.; Takeuchi, T.; Kohyama-Koganeya, A.; Nakajima, K.; Hirabayashi, Y.; Wada, K.; Nagai, Y. Glucocerebrosidase deficiency accelerates the accumulation of proteinase K-resistant α-synuclein and aggravates neurodegeneration in a Drosophila model of Parkinson’s disease. Hum. Mol. Genet. 2015, 24, 6675–6686. [Google Scholar] [CrossRef] [PubMed]
  29. Choong, C.J.; Aguirre, C.; Kakuda, K.; Beck, G.; Nakanishi, H.; Kimura, Y.; Shimma, S.; Nabekura, K.; Hideshima, M.; Doi, J.; et al. Phosphatidylinositol-3,4,5-trisphosphate interacts with alpha-synuclein and initiates its aggregation and formation of Parkinson’s disease-related fibril polymorphism. Acta Neuropathol. 2023, 145, 573–595. [Google Scholar] [CrossRef]
  30. Morgan, N.V.; Westaway, S.K.; Morton, J.E.; Gregory, A.; Gissen, P.; Sonek, S.; Cangul, H.; Coryell, J.; Canham, N.; Nardocci, N.; et al. PLA2G6, encoding a phospholipase A2, is mutated in neurodegenerative disorders with high brain iron. Nat. Genet. 2006, 38, 752–754. [Google Scholar] [CrossRef]
  31. Darvish, H.; Bravo, P.; Tafakhori, A.; Azcona, L.J.; Ranji-Burachaloo, S.; Johari, A.H.; Paisán-Ruiz, C. Identification of a large homozygous VPS13C deletion in a patient with early-onset Parkinsonism. Mov. Disord. 2018, 33, 1968–1970. [Google Scholar] [CrossRef] [PubMed]
  32. Illés, A.; Csabán, D.; Grosz, Z.; Balicza, P.; Gézsi, A.; Molnár, V.; Bencsik, R.; Gál, A.; Klivényi, P.; Molnar, M.J. The Role of Genetic Testing in the Clinical Practice and Research of Early-Onset Parkinsonian Disorders in a Hungarian Cohort: Increasing Challenge in Genetic Counselling, Improving Chances in Stratification for Clinical Trials. Front. Genet. 2019, 10, 1061. [Google Scholar] [CrossRef] [PubMed]
  33. Marshall, M.S.; Bongarzone, E.R. Beyond Krabbe’s disease: The potential contribution of galactosylceramidase deficiency to neuronal vulnerability in late-onset synucleinopathies. J. Neurosci. Res. 2016, 94, 1328–1332. [Google Scholar] [CrossRef] [PubMed]
  34. Araki, K.; Yagi, N.; Ikemoto, Y.; Yagi, H.; Choong, C.J.; Hayakawa, H.; Beck, G.; Sumi, H.; Fujimura, H.; Moriwaki, T.; et al. Synchrotron FTIR micro-spectroscopy for structural analysis of Lewy bodies in the brain of Parkinson’s disease patients. Sci. Rep. 2015, 5, 17625. [Google Scholar] [CrossRef]
  35. Lavedan, C.; Buchholtz, S.; Auburger, G.; Albin, R.L.; Athanassiadou, A.; Blancato, J.; Burguera, J.A.; Ferrell, R.E.; Kostic, V.; Leroy, E.; et al. Absence of mutation in the beta- and gamma-synuclein genes in familial autosomal dominant Parkinson’s disease. DNA Res. 1998, 5, 401–402. [Google Scholar] [CrossRef] [PubMed]
  36. George, J.M.; Jin, H.; Woods, W.S.; Clayton, D.F. Characterization of a novel protein regulated during the critical period for song learning in the zebra finch. Neuron 1995, 15, 361–372. [Google Scholar] [CrossRef] [PubMed]
  37. Iwai, A.; Masliah, E.; Yoshimoto, M.; Ge, N.; Flanagan, L.; de Silva, H.A.; Kittel, A.; Saitoh, T. The precursor protein of non-A beta component of Alzheimer’s disease amyloid is a presynaptic protein of the central nervous system. Neuron 1995, 14, 467–475. [Google Scholar] [CrossRef]
  38. Maroteaux, L.; Campanelli, J.T.; Scheller, R.H. Synuclein: A neuron-specific protein localized to the nucleus and presynaptic nerve terminal. J. Neurosci. 1988, 8, 2804–2815. [Google Scholar] [CrossRef] [PubMed]
  39. Withers, G.S.; George, J.M.; Banker, G.A.; Clayton, D.F. Delayed localization of synelfin (synuclein, NACP) to presynaptic terminals in cultured rat hippocampal neurons. Brain Res. Dev. Brain Res. 1997, 99, 87–94. [Google Scholar] [CrossRef]
  40. Jensen, P.H.; Nielsen, M.S.; Jakes, R.; Dotti, C.G.; Goedert, M. Binding of alpha-synuclein to brain vesicles is abolished by familial Parkinson’s disease mutation. J. Biol. Chem. 1998, 273, 26292–26294. [Google Scholar] [CrossRef]
  41. Fortin, D.L.; Troyer, M.D.; Nakamura, K.; Kubo, S.; Anthony, M.D.; Edwards, R.H. Lipid rafts mediate the synaptic localization of alpha-synuclein. J. Neurosci. 2004, 24, 6715–6723. [Google Scholar] [CrossRef]
  42. Kahle, P.J.; Neumann, M.; Ozmen, L.; Muller, V.; Jacobsen, H.; Schindzielorz, A.; Okochi, M.; Leimer, U.; van Der Putten, H.; Probst, A.; et al. Subcellular localization of wild-type and Parkinson’s disease-associated mutant alpha -synuclein in human and transgenic mouse brain. J. Neurosci. 2000, 20, 6365–6373. [Google Scholar] [CrossRef]
  43. Lee, H.J.; Choi, C.; Lee, S.J. Membrane-bound alpha-synuclein has a high aggregation propensity and the ability to seed the aggregation of the cytosolic form. J. Biol. Chem. 2002, 277, 671–678. [Google Scholar] [CrossRef]
  44. Jakes, R.; Spillantini, M.G.; Goedert, M. Identification of two distinct synucleins from human brain. FEBS Lett. 1994, 345, 27–32. [Google Scholar] [CrossRef]
  45. Weinreb, P.H.; Zhen, W.; Poon, A.W.; Conway, K.A.; Lansbury, P.T., Jr. NACP, a protein implicated in Alzheimer’s disease and learning, is natively unfolded. Biochemistry 1996, 35, 13709–13715. [Google Scholar] [CrossRef] [PubMed]
  46. Zhu, M.; Li, J.; Fink, A.L. The association of α-synuclein with membranes affects bilayer structure, stability, and fibril formation. J. Biol. Chem. 2003, 278, 40186–40197. [Google Scholar] [CrossRef]
  47. Zhang, Z.; Kang, S.S.; Liu, X.; Ahn, E.H.; Zhang, Z.; He, L.; Iuvone, P.M.; Duong, D.M.; Seyfried, N.T.; Benskey, M.J.; et al. Asparagine endopeptidase cleaves α-synuclein and mediates pathologic activities in Parkinson’s disease. Nat. Struct. Mol. Biol. 2017, 24, 632–642. [Google Scholar] [CrossRef]
  48. Gai, W.P.; Yuan, H.X.; Li, X.Q.; Power, J.T.; Blumbergs, P.C.; Jensen, P.H. In situ and in vitro study of colocalization and segregation of α-synuclein, ubiquitin, and lipids in Lewy bodies. Exp. Neurol. 2000, 166, 324–333. [Google Scholar] [CrossRef]
  49. Halliday, G.M.; Ophof, A.; Broe, M.; Jensen, P.H.; Kettle, E.; Fedorow, H.; Cartwright, M.I.; Griffiths, F.M.; Shepherd, C.E.; Double, K.L. α-synuclein redistributes to neuromelanin lipid in the substantia nigra early in Parkinson’s disease. Brain 2005, 128, 2654–2664. [Google Scholar] [CrossRef]
  50. Giasson, B.I.; Murray, I.V.; Trojanowski, J.Q.; Lee, V.M. A hydrophobic stretch of 12 amino acid residues in the middle of α-synuclein is essential for filament assembly. J. Biol. Chem. 2001, 276, 2380–2386. [Google Scholar] [CrossRef] [PubMed]
  51. Periquet, M.; Fulga, T.; Myllykangas, L.; Schlossmacher, M.G.; Feany, M.B. Aggregated α-synuclein mediates dopaminergic neurotoxicity in vivo. J. Neurosci. 2007, 27, 3338–3346. [Google Scholar] [CrossRef] [PubMed]
  52. Rodriguez, J.A.; Ivanova, M.I.; Sawaya, M.R.; Cascio, D.; Reyes, F.E.; Shi, D.; Sangwan, S.; Guenther, E.L.; Johnson, L.M.; Zhang, M.; et al. Structure of the toxic core of α-synuclein from invisible crystals. Nature 2015, 525, 486–490. [Google Scholar] [CrossRef] [PubMed]
  53. Bodles, A.M.; Guthrie, D.J.; Greer, B.; Irvine, G.B. Identification of the region of non-Aβ component (NAC) of Alzheimer’s disease amyloid responsible for its aggregation and toxicity. J. Neurochem. 2001, 78, 384–395. [Google Scholar] [CrossRef]
  54. Hong, D.P.; Xiong, W.; Chang, J.Y.; Jiang, C. The role of the C-terminus of human α-synuclein: Intra-disulfide bonds between the C-terminus and other regions stabilize non-fibrillar monomeric isomers. FEBS Lett. 2011, 585, 561–566. [Google Scholar] [CrossRef] [PubMed]
  55. Bertoncini, C.W.; Jung, Y.S.; Fernandez, C.O.; Hoyer, W.; Griesinger, C.; Jovin, T.M.; Zweckstetter, M. Release of long-range tertiary interactions potentiates aggregation of natively unstructured α-synuclein. Proc. Natl. Acad. Sci. USA 2005, 102, 1430–1435. [Google Scholar] [CrossRef] [PubMed]
  56. Calo, L.; Wegrzynowicz, M.; Santivanez-Perez, J.; Grazia Spillantini, M. Synaptic failure and alpha-synuclein. Mov. Disord. 2016, 31, 169–177. [Google Scholar] [CrossRef] [PubMed]
  57. Larsen, K.E.; Schmitz, Y.; Troyer, M.D.; Mosharov, E.; Dietrich, P.; Quazi, A.Z.; Savalle, M.; Nemani, V.; Chaudhry, F.A.; Edwards, R.H.; et al. Alpha-synuclein overexpression in PC12 and chromaffin cells impairs catecholamine release by interfering with a late step in exocytosis. J. Neurosci. 2006, 26, 11915–11922. [Google Scholar] [CrossRef] [PubMed]
  58. Nemani, V.M.; Lu, W.; Berge, V.; Nakamura, K.; Onoa, B.; Lee, M.K.; Chaudhry, F.A.; Nicoll, R.A.; Edwards, R.H. Increased expression of alpha-synuclein reduces neurotransmitter release by inhibiting synaptic vesicle reclustering after endocytosis. Neuron 2010, 65, 66–79. [Google Scholar] [CrossRef] [PubMed]
  59. Lundblad, M.; Decressac, M.; Mattsson, B.; Bjorklund, A. Impaired neurotransmission caused by overexpression of alpha-synuclein in nigral dopamine neurons. Proc. Natl. Acad. Sci. USA 2012, 109, 3213–3219. [Google Scholar] [CrossRef]
  60. Phan, J.A.; Stokholm, K.; Zareba-Paslawska, J.; Jakobsen, S.; Vang, K.; Gjedde, A.; Landau, A.M.; Romero-Ramos, M. Early synaptic dysfunction induced by alpha-synuclein in a rat model of Parkinson’s disease. Sci. Rep. 2017, 7, 6363. [Google Scholar] [CrossRef]
  61. Bussell, R., Jr.; Eliezer, D. Effects of Parkinson’s disease-linked mutations on the structure of lipid-associated alpha-synuclein. Biochemistry 2004, 43, 4810–4818. [Google Scholar] [CrossRef]
  62. Fredenburg, R.A.; Rospigliosi, C.; Meray, R.K.; Kessler, J.C.; Lashuel, H.A.; Eliezer, D.; Lansbury, P.T., Jr. The impact of the E46K mutation on the properties of alpha-synuclein in its monomeric and oligomeric states. Biochemistry 2007, 46, 7107–7118. [Google Scholar] [CrossRef] [PubMed]
  63. Burre, J.; Sharma, M.; Tsetsenis, T.; Buchman, V.; Etherton, M.R.; Sudhof, T.C. Alpha-synuclein promotes SNARE-complex assembly in vivo and in vitro. Science 2010, 329, 1663–1667. [Google Scholar] [CrossRef] [PubMed]
  64. Diao, J.; Burre, J.; Vivona, S.; Cipriano, D.J.; Sharma, M.; Kyoung, M.; Sudhof, T.C.; Brunger, A.T. Native alpha-synuclein induces clustering of synaptic-vesicle mimics via binding to phospholipids and synaptobrevin-2/VAMP2. Elife 2013, 2, e00592. [Google Scholar] [CrossRef] [PubMed]
  65. Logan, T.; Bendor, J.; Toupin, C.; Thorn, K.; Edwards, R.H. alpha-Synuclein promotes dilation of the exocytotic fusion pore. Nat. Neurosci. 2017, 20, 681–689. [Google Scholar] [CrossRef] [PubMed]
  66. Wang, C.; Zhang, K.; Cai, B.; Haller, J.E.; Carnazza, K.E.; Hu, J.; Zhao, C.; Tian, Z.; Hu, X.; Hall, D.; et al. VAMP2 chaperones α-synuclein in synaptic vesicle co-condensates. Nat. Cell Biol. 2024, 26, 1287–1295. [Google Scholar] [CrossRef]
  67. Davidson, W.S.; Jonas, A.; Clayton, D.F.; George, J.M. Stabilization of alpha-synuclein secondary structure upon binding to synthetic membranes. J. Biol. Chem. 1998, 273, 9443–9449. [Google Scholar] [CrossRef]
  68. Mori, A.; Hatano, T.; Inoshita, T.; Shiba-Fukushima, K.; Koinuma, T.; Meng, H.; Kubo, S.I.; Spratt, S.; Cui, C.; Yamashita, C.; et al. Parkinson’s disease-associated iPLA2-VIA/PLA2G6 regulates neuronal functions and alpha-synuclein stability through membrane remodeling. Proc. Natl. Acad. Sci. USA 2019, 116, 20689–20699. [Google Scholar] [CrossRef] [PubMed]
  69. Perrin, R.J.; Woods, W.S.; Clayton, D.F.; George, J.M. Interaction of human alpha-Synuclein and Parkinson’s disease variants with phospholipids. Structural analysis using site-directed mutagenesis. J. Biol. Chem. 2000, 275, 34393–34398. [Google Scholar] [CrossRef]
  70. Rhoades, E.; Ramlall, T.F.; Webb, W.W.; Eliezer, D. Quantification of alpha-synuclein binding to lipid vesicles using fluorescence correlation spectroscopy. Biophys. J. 2006, 90, 4692–4700. [Google Scholar] [CrossRef]
  71. Galvagnion, C.; Brown, J.W.; Ouberai, M.M.; Flagmeier, P.; Vendruscolo, M.; Buell, A.K.; Sparr, E.; Dobson, C.M. Chemical properties of lipids strongly affect the kinetics of the membrane-induced aggregation of alpha-synuclein. Proc. Natl. Acad. Sci. USA 2016, 113, 7065–7070. [Google Scholar] [CrossRef] [PubMed]
  72. Kubo, S.; Nemani, V.M.; Chalkley, R.J.; Anthony, M.D.; Hattori, N.; Mizuno, Y.; Edwards, R.H.; Fortin, D.L. A combinatorial code for the interaction of alpha-synuclein with membranes. J. Biol. Chem. 2005, 280, 31664–31672. [Google Scholar] [CrossRef] [PubMed]
  73. Zakharova, I.O.; Sokolova, T.V.; Furaev, V.V.; Rychkova, M.P.; Avrova, N.F. Effects of oxidative stress inducers, neurotoxins, and ganglioside GM1 on Na+, K+-ATPase in PC12 and brain synaptosomes. Zhurnal Evoliutsionnoi Biokhimii Fiziol. 2007, 43, 148–154. [Google Scholar]
  74. Middleton, E.R.; Rhoades, E. Effects of curvature and composition on alpha-synuclein binding to lipid vesicles. Biophys. J. 2010, 99, 2279–2288. [Google Scholar] [CrossRef] [PubMed]
  75. van Meer, G.; Voelker, D.R.; Feigenson, G.W. Membrane lipids: Where they are and how they behave. Nat. Rev. Mol. Cell Biol. 2008, 9, 112–124. [Google Scholar] [CrossRef]
  76. Kobayashi, T.; Beuchat, M.H.; Chevallier, J.; Makino, A.; Mayran, N.; Escola, J.M.; Lebrand, C.; Cosson, P.; Kobayashi, T.; Gruenberg, J. Separation and characterization of late endosomal membrane domains. J. Biol. Chem. 2002, 277, 32157–32164. [Google Scholar] [CrossRef] [PubMed]
  77. O’Leary, E.I.; Jiang, Z.; Strub, M.P.; Lee, J.C. Effects of phosphatidylcholine membrane fluidity on theconformation and aggregation of N-terminally acetylated alpha-synuclein. J. Biol. Chem. 2018, 293, 11195–11205. [Google Scholar] [CrossRef]
  78. Chandra, S.; Chen, X.; Rizo, J.; Jahn, R.; Sudhof, T.C. A broken alpha -helix in folded alpha -Synuclein. J. Biol. Chem. 2003, 278, 15313–15318. [Google Scholar] [CrossRef] [PubMed]
  79. Drescher, M.; Veldhuis, G.; van Rooijen, B.D.; Milikisyants, S.; Subramaniam, V.; Huber, M. Antiparallel arrangement of the helices of vesicle-bound alpha-synuclein. J. Am. Chem. Soc. 2008, 130, 7796–7797. [Google Scholar] [CrossRef]
  80. Bodner, C.R.; Dobson, C.M.; Bax, A. Multiple tight phospholipid-binding modes of alpha-synuclein revealed by solution NMR spectroscopy. J. Mol. Biol. 2009, 390, 775–790. [Google Scholar] [CrossRef]
  81. Irwin, D.J.; Lee, V.M.; Trojanowski, J.Q. Parkinson’s disease dementia: Convergence of alpha-synuclein, tau and amyloid-beta pathologies. Nat. Rev. Neurosci. 2013, 14, 626–636. [Google Scholar] [CrossRef]
  82. Kordower, J.H.; Chu, Y.; Hauser, R.A.; Freeman, T.B.; Olanow, C.W. Lewybody-likepathologyinlong-term embryonic nigral transplants in Parkinson’s disease. Nat. Med. 2008, 14, 504–506. [Google Scholar] [CrossRef] [PubMed]
  83. Kordower, J.H.; Chu, Y.; Hauser, R.A.; Olanow, C.W.; Freeman, T.B. Transplanteddopaminergicneurons develop PD pathologic changes: A second case report. Mov. Disord. 2008, 23, 2303–2306. [Google Scholar] [CrossRef]
  84. Li, J.Y.; Englund, E.; Holton, J.L.; Soulet, D.; Hagell, P.; Lees, A.J.; Lashley, T.; Quinn, N.P.; Rehncrona, S.; Bjorklund, A.; et al. Lewy bodies in grafted neurons in subjects with Parkinson’s disease suggest host-to-graft disease propagation. Nat. Med. 2008, 14, 501–503. [Google Scholar] [CrossRef]
  85. 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]
  86. Hansen, C.; Angot, E.; Bergstrom, A.L.; Steiner, J.A.; Pieri, L.; Paul, G.; Outeiro, T.F.; Melki, R.; Kallunki, P.; Fog, K.; et al. alpha-Synuclein propagates from mouse brain to grafted dopaminergic neurons and seeds aggregation in cultured human cells. J. Clin. Investig. 2011, 121, 715–725. [Google Scholar] [CrossRef]
  87. Kordower, J.H.; Dodiya, H.B.; Kordower, A.M.; Terpstra, B.; Paumier, K.; Madhavan, L.; Sortwell, C.; Steece-Collier, K.; Collier, T.J. Transfer of host-derived alpha synuclein to grafted dopaminergic neurons in rat. Neurobiol. Dis. 2011, 43, 552–557. [Google Scholar] [CrossRef] [PubMed]
  88. Luk, K.C.; Kehm, V.; Carroll, J.; Zhang, B.; O’Brien, P.; Trojanowski, J.Q.; Lee, V.M. Pathological alpha-synuclein transmission initiates Parkinson-like neurodegeneration in nontransgenic mice. Science 2012, 338, 949–953. [Google Scholar] [CrossRef] [PubMed]
  89. Shimozawa, A.; Ono, M.; Takahara, D.; Tarutani, A.; Imura, S.; Masuda-Suzukake, M.; Higuchi, M.; Yanai, K.; Hisanaga, S.I.; Hasegawa, M. Propagation of pathological alpha-synuclein in marmoset brain. Acta Neuropathol. Commun. 2017, 5, 12. [Google Scholar] [CrossRef]
  90. Okuzumi, A.; Kurosawa, M.; Hatano, T.; Takanashi, M.; Nojiri, S.; Fukuhara, T.; Yamanaka, T.; Miyazaki, H.; Yoshinaga, S.; Furukawa, Y.; et al. Rapid dissemination of alpha-synuclein seeds through neural circuits in an in-vivo prion-like seeding experiment. Acta Neuropathol. Commun. 2018, 6, 96. [Google Scholar] [CrossRef]
  91. Jo, E.; McLaurin, J.; Yip, C.M.; St George-Hyslop, P.; Fraser, P.E. alpha-Synuclein membrane interactions and lipid specificity. J. Biol. Chem. 2000, 275, 34328–34334. [Google Scholar] [CrossRef] [PubMed]
  92. Fares, M.B.; Ait-Bouziad, N.; Dikiy, I.; Mbefo, M.K.; Jovicic, A.; Kiely, A.; Holton, J.L.; Lee, S.J.; Gitler, A.D.; Eliezer, D.; et al. The novel Parkinson’s disease linked mutation G51D attenuates in vitro aggregation and membrane binding of alpha-synuclein, and enhances its secretion and nuclear localization in cells. Hum. Mol. Genet. 2014, 23, 4491–4509. [Google Scholar] [CrossRef] [PubMed]
  93. Bodner, C.R.; Maltsev, A.S.; Dobson, C.M.; Bax, A. Differential phospholipid binding of alpha-synuclein variants implicated in Parkinson’s disease revealed by solution NMR spectroscopy. Biochemistry 2010, 49, 862–871. [Google Scholar] [CrossRef] [PubMed]
  94. Khalaf, O.; Fauvet, B.; Oueslati, A.; Dikiy, I.; Mahul-Mellier, A.L.; Ruggeri, F.S.; Mbefo, M.K.; Vercruysse, F.; Dietler, G.; Lee, S.J.; et al. The H50Q mutation enhances alpha-synuclein aggregation, secretion, and toxicity. J. Biol. Chem. 2014, 289, 21856–21876. [Google Scholar] [CrossRef] [PubMed]
  95. Choi, W.; Zibaee, S.; Jakes, R.; Serpell, L.C.; Davletov, B.; Crowther, R.A.; Goedert, M. Mutation E46K increases phospholipid binding and assembly into filaments of human alpha-synuclein. FEBS Lett. 2004, 576, 363–368. [Google Scholar] [CrossRef] [PubMed]
  96. 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]
  97. Ghosh, D.; Mondal, M.; Mohite, G.M.; Singh, P.K.; Ranjan, P.; Anoop, A.; Ghosh, S.; Jha, N.N.; Kumar, A.; Maji, S.K. The Parkinson’s disease-associated H50Q mutation accelerates α-Synuclein aggregation in vitro. Biochemistry 2013, 52, 6925–6927. [Google Scholar] [CrossRef]
  98. Shahmoradian, S.H.; Lewis, A.J.; Genoud, C.; Hench, J.; Moors, T.E.; Navarro, P.P.; Castano-Diez, D.; Schweighauser, G.; Graff-Meyer, A.; Goldie, K.N.; et al. Lewy pathology in Parkinson’s disease consists of crowded organelles and lipid membranes. Nat. Neurosci. 2019, 22, 1099–1109. [Google Scholar] [CrossRef]
  99. 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]
  100. Volles, M.J.; Lee, S.J.; Rochet, J.C.; Shtilerman, M.D.; Ding, T.T.; Kessler, J.C.; Lansbury, P.T., Jr. Vesicle permeabilization by protofibrillar alpha-synuclein: Implications for the pathogenesis and treatment of Parkinson’s disease. Biochemistry 2001, 40, 7812–7819. [Google Scholar] [CrossRef]
  101. Volles, M.J.; Lansbury, P.T., Jr. Vesicle permeabilization by protofibrillar alpha-synuclein is sensitive to Parkinson’s disease-linked mutations and occurs by a pore-like mechanism. Biochemistry 2002, 41, 4595–4602. [Google Scholar] [CrossRef] [PubMed]
  102. Stefanovic, A.N.; Lindhoud, S.; Semerdzhiev, S.A.; Claessens, M.M.; Subramaniam, V. Oligomers of Parkinson’s Disease-Related α-Synuclein Mutants Have Similar Structures but Distinctive Membrane Permeabilization Properties. Biochemistry 2015, 54, 3142–3150. [Google Scholar] [CrossRef]
  103. Uversky, V.N.; Li, J.; Fink, A.L. Metal-triggered structural transformations, aggregation, and fibrillation of human α-synuclein. A possible molecular NK between Parkinson’s disease and heavy metal exposure. J. Biol. Chem. 2001, 276, 44284–44296. [Google Scholar] [CrossRef] [PubMed]
  104. Kahle, P.J.; Neumann, M.; Ozmen, L.; Haass, C. 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. 2000, 920, 33–41. [Google Scholar] [CrossRef] [PubMed]
  105. Huebbe, P.; Rimbach, G. Evolution of human apolipoprotein E (APOE) isoforms: Gene structure, protein function and interaction with dietary factors. Ageing Res. Rev. 2017, 37, 146–161. [Google Scholar] [CrossRef] [PubMed]
  106. Varkey, J.; Isas, J.M.; Mizuno, N.; Jensen, M.B.; Bhatia, V.K.; Jao, C.C.; Petrlova, J.; Voss, J.C.; Stamou, D.G.; Steven, A.C.; et al. Membrane curvature induction and tubulation are common features of synucleins and apolipoproteins. J. Biol. Chem. 2010, 285, 32486–32493. [Google Scholar] [CrossRef] [PubMed]
  107. Elliott, D.A.; Weickert, C.S.; Garner, B. Apolipoproteins in the brain: Implications for neurological and psychiatric disorders. Clin. Lipidol. 2010, 51, 555–573. [Google Scholar] [CrossRef] [PubMed]
  108. Tamam, Y.; Tasdemir, N.; Yalman, M.; Tamam, B. Association of apolipoprotein E genotypes with prognosis in multiple sclerosis. Eur. Rev. Med. Pharmacol. Sci. 2011, 15, 1122–1130. [Google Scholar] [PubMed]
  109. Ponsford, J.; McLaren, A.; Schönberger, M.; Burke, R.; Rudzki, D.; Olver, J.; Ponsford, M. The association between apolipoprotein E and traumatic brain injury severity and functional outcome in a rehabilitation sample. J. Neurotrauma 2011, 28, 1683–1692. [Google Scholar] [CrossRef]
  110. Amouyel, P.; Vidal, O.; Launay, J.M.; Laplanche, J.L. The apolipoprotein E alleles as major susceptibility factors for Creutzfeldt-Jakob disease. The French Research Group on Epidemiology of Human Spongiform Encephalopathies. Lancet 1994, 344, 1315–1318. [Google Scholar] [CrossRef]
  111. Wilhelmus, M.M.; Bol, J.G.; Van Haastert, E.S.; Rozemuller, A.J.; Bu, G.; Drukarch, B.; Hoozemans, J.J. Apolipoprotein E and LRP1 increase early in Parkinson’s disease pathogenesis. Am. J. Pathol. 2011, 179, 2152–2156. [Google Scholar] [CrossRef] [PubMed]
  112. Emamzadeh, F.N.; Aojula, H.; McHugh, P.C.; Allsop, D. Effects of different isoforms of apoE on aggregation of the α-synuclein protein implicated in Parkinson’s disease. Neurosci. Lett. 2016, 618, 146–151. [Google Scholar] [CrossRef] [PubMed]
  113. Paslawski, W.; Zareba-Paslawska, J.; Zhang, X.; Hölzl, K.; Wadensten, H.; Shariatgorji, M.; Janelidze, S.; Hansson, O.; Forsgren, L.; Andrén, P.E.; et al. α-synuclein-lipoprotein interactions and elevated ApoE level in cerebrospinal fluid from Parkinson’s disease patients. Proc. Natl. Acad. Sci. USA 2019, 116, 15226–15235. [Google Scholar] [CrossRef] [PubMed]
  114. Brangwynne, C.P.; Eckmann, C.R.; Courson, D.S.; Rybarska, A.; Hoege, C.; Gharakhani, J.; Jülicher, F.; Hyman, A.A. Germline P granules are liquid droplets that localize by controlled dissolution/condensation. Science 2009, 324, 1729–1732. [Google Scholar] [CrossRef]
  115. Hyman, A.A.; Weber, C.A.; Jülicher, F. Liquid-liquid phase separation in biology. Annu. Rev. Cell Dev. Biol. 2014, 30, 39–58. [Google Scholar] [CrossRef] [PubMed]
  116. Banani, S.F.; Rice, A.M.; Peeples, W.B.; Lin, Y.; Jain, S.; Parker, R.; Rosen, M.K. Compositional control of phase-separated cellular bodies. Cell 2016, 166, 651–663. [Google Scholar] [CrossRef] [PubMed]
  117. Shin, Y.; Brangwynne, C.P. Liquid phase condensation in cell physiology and disease. Science 2017, 357, eaaf4382. [Google Scholar] [CrossRef] [PubMed]
  118. Alberti, S. Phase separation in biology. Curr. Biol. 2017, 27, R1097–R1102. [Google Scholar] [CrossRef]
  119. Boeynaems, S.; Alberti, S.; Fawzi, N.L.; Mittag, T.; Polymenidou, M.; Rousseau, F.; Schymkowitz, J.; Shorter, J.; Wolozin, B.; Van Den Bosch, L.; et al. Protein phase separation: A new phase in cell biology. Trends Cell Biol. 2018, 28, 420–435. [Google Scholar] [CrossRef]
  120. Holehouse, A.S.; Pappu, R.V. Functional Implications of Intracellular Phase Transitions. Biochemistry 2018, 57, 2415–2423. [Google Scholar] [CrossRef]
  121. Patel, A.; Lee, H.O.; Jawerth, L.; Maharana, S.; Jahnel, M.; Hein, M.Y.; Stoynov, S.; Mahamid, J.; Saha, S.; Franzmann, T.M.; et al. A liquid-to-solid phase transition of the ALS protein FUS accelerated by disease mutation. Cell 2015, 162, 1066–1077. [Google Scholar] [CrossRef]
  122. Conicella, A.E.; Zerze, G.H.; Mittal, J.; Fawzi, N.L. ALS mutations disrupt phase separation mediated by α-helical structure in the TDP-43 low-complexity C-terminal domain. Structure 2016, 24, 1537–1549. [Google Scholar] [CrossRef]
  123. Wegmann, S.; Eftekharzadeh, B.; Tepper, K.; Zoltowska, K.M.; Bennett, R.E.; Dujardin, S.; Laskowski, P.R.; MacKenzie, D.; Kamath, T.; Commins, C.; et al. Tau protein liquid-liquid phase separation can initiate tau aggregation. EMBO J. 2018, 37, e98049. [Google Scholar] [CrossRef]
  124. Ambadipudi, S.; Biernat, J.; Riedel, D.; Mandelkow, E.; Zweckstetter, M. Liquid-liquid phase separation of the microtubule-binding repeats of the Alzheimer-related protein Tau. Nat. Commun. 2017, 8, 275. [Google Scholar] [CrossRef]
  125. Molliex, A.; Temirov, J.; Lee, J.; Coughlin, M.; Kanagaraj, A.P.; Kim, H.J.; Mittag, T.; Taylor, J.P. Phase separation by low complexity domains promotes stress granule assembly and drives pathological fibrillization. Cell 2015, 163, 123–133. [Google Scholar] [CrossRef]
  126. Ray, S.; Singh, N.; Kumar, R.; Patel, K.; Pandey, S.; Datta, D.; Mahato, J.; Panigrahi, R.; Navalkar, A.; Mehra, S.; et al. α-Synuclein aggregation nucleates through liquid-liquid phase separation. Nat. Chem. 2020, 12, 705–716. [Google Scholar] [CrossRef]
  127. Huang, S.; Xu, B.; Liu, Y. Calcium promotes α-synuclein liquid-liquid phase separation to accelerate amyloid aggregation. Biochem. Biophys. Res. Commun. 2022, 603, 13–20. [Google Scholar] [CrossRef]
  128. Van Maarschalkerweerd, A.; Vetri, V.; Langkilde, A.E.; Foderà, V.; Vestergaard, B. Protein/lipid coaggregates are formed during α-synuclein-induced disruption of lipid bilayers. Biomacromolecules 2014, 15, 3643–3654. [Google Scholar] [CrossRef] [PubMed]
  129. Roy, S.; Wolman, L. Ultrastructural observations in Parkinsonism. J. Pathol. 1969, 99, 39–44. [Google Scholar] [CrossRef] [PubMed]
  130. Kosaka, K.; Oyanagi, S.; Matsushita, M.; Hori, A. Presenile dementia with Alzheimer-, Pick- and Lewy-body changes. Acta Neuropathol. 1976, 36, 221–233. [Google Scholar] [CrossRef] [PubMed]
  131. Trojanowski, J.Q.; Lee, V.M. Aggregation of neurofilament and alpha-synuclein proteins in Lewy bodies: Implications for the pathogenesis of Parkinson disease and Lewy body dementia. Arch. Neurol. 1998, 55, 151–152. [Google Scholar] [CrossRef]
  132. Jensen, P.H.; Islam, K.; Kenney, J.; Nielsen, M.S.; Power, J.; Gai, W.P. Microtubule-associated protein 1B is a component of cortical Lewy bodies and binds alpha-synuclein filaments. J. Biol. Chem. 2000, 275, 21500–21507. [Google Scholar] [CrossRef] [PubMed]
  133. Flavin, W.P.; Bousset, L.; Green, Z.C.; Chu, Y.; Skarpathiotis, S.; Chaney, M.J.; Kordower, J.H.; Melki, R.; Campbell, E.M. Endocytic vesicle rupture is a conserved mechanism of cellular invasion by amyloid proteins. Acta Neuropathol. 2017, 134, 629–653. [Google Scholar] [CrossRef] [PubMed]
  134. den Jager, W.A. Sphingomyelin in Lewy inclusion bodies in Parkinson’s disease. Arch. Neurol. 1969, 21, 615–619. [Google Scholar] [CrossRef] [PubMed]
  135. Issidorides, M.R.; Mytilineou, C.; Panayotacopoulou, M.T.; Yahr, M.D. Lewy bodies in parkinsonism share components with intraneuronal protein bodies of normal brains. J. Neural. Transm. Park. Dis. Dement. Sect. 1991, 3, 49–61. [Google Scholar] [CrossRef] [PubMed]
  136. Sidransky, E.; Nalls, M.A.; Aasly, J.O.; Aharon-Peretz, J.; Annesi, G.; Barbosa, E.R.; Bar-Shira, A.; Berg, D.; Bras, J.; Brice, A.; et al. Multicenter analysis of glucocerebrosidase mutations in Parkinson’s disease. N. Engl. J. Med. 2009, 361, 1651–1661. [Google Scholar] [CrossRef] [PubMed]
  137. Migdalska-Richards, A.; Schapira, A.H. The relationship between glucocerebrosidase mutations and Parkinson disease. J. Neurochem. 2016, 139, 77–90. [Google Scholar] [CrossRef] [PubMed]
  138. Mizuno, H.; Fujikake, N.; Wada, K.; Nagai, Y. α-Synuclein transgenic Drosophila as a model of Parkinson’s disease and related synucleinopathies. Park. Dis. 2010, 2011, 212706. [Google Scholar] [CrossRef]
  139. Zunke, F.; Moise, A.C.; Belur, N.R.; Gelyana, E.; Stojkovska, I.; Dzaferbegovic, H.; Toker, N.J.; Jeon, S.; Fredriksen, K.; Mazzulli, J.R. Reversible conformational conversion of α-synuclein into toxic assemblies by glucosylceramide. Neuron 2018, 97, 92–107. [Google Scholar] [CrossRef]
  140. Shachar, T.; Lo Bianco, C.; Recchia, A.; Wiessner, C.; Raas-Rothschild, A.; Futerman, A.H. Lysosomal storage disorders and Parkinson’s disease: Gaucher disease and beyond. Mov. Disord. 2011, 26, 1593–1604. [Google Scholar] [CrossRef]
  141. Walkley, S.U. Secondary accumulation of gangliosides in lysosomal storage disorders. Semin. Cell Dev. Biol. 2004, 15, 433–444. [Google Scholar] [CrossRef] [PubMed]
  142. Gegg, M.E.; Burke, D.; Heales, S.J.; Cooper, J.M.; Hardy, J.; Wood, N.W.; Schapira, A.H.V. Glucocerebrosidase deficiency in substantia nigra of parkinson disease brains. Ann. Neurol. 2012, 72, 455–463. [Google Scholar] [CrossRef] [PubMed]
  143. Moors, T.E.; Paciotti, S.; Ingrassia, A.; Quadri, M.; Breedveld, G.; Tasegian, A.; Chiasserini, D.; Eusebi, P.; Duran-Pacheco, G.; Kremer, T.; et al. Characterization of brain lysosomal activities in GBA-related and sporadic Parkinson’s disease and dementia with Lewy bodies. Mol. Neurobiol. 2018, 56, 1344–1355. [Google Scholar] [CrossRef]
  144. Murphy, K.E.; Gysbers, A.M.; Abbott, S.K.; Tayebi, N.; Kim, W.S.; Sidransky, E.; Cooper, A.; Garner, B.; Halliday, G.M. Reduced glucocerebrosidase is associated with increased alpha- synuclein in sporadic Parkinson’s disease. Brain 2014, 137, 834–848. [Google Scholar] [CrossRef]
  145. Chiasserini, D.; Paciotti, S.; Eusebi, P.; Persichetti, E.; Tasegian, A.; Kurzawa-Akanbi, M.; Chinnery, P.F.; Morris, C.M.; Calabresi, P.; Parnetti, L.; et al. Selective loss of glucocerebrosidase activity in sporadic Parkinson’s disease and dementia with Lewy bodies. Mol. Neurodegener. 2015, 10, 15. [Google Scholar] [CrossRef]
  146. Quadri, M.; Fang, M.; Picillo, M.; Olgiati, S.; Breedveld, G.J.; Graafland, J.; Wu, B.; Xu, F.; Erro, R.; Amboni, M.; et al. Mutation in the SYNJ1 gene associated with autosomal recessive, early-onset Parkinsonism. Hum. Mutat. 2013, 34, 1208–1215. [Google Scholar] [CrossRef] [PubMed]
  147. Sumi-Akamaru, H.; Beck, G.; Shinzawa, K.; Kato, S.; Riku, Y.; Yoshida, M.; Fujimura, H.; Tsujimoto, Y.; Sakoda, S.; Mochizuki, H. High expression of α-synuclein in damaged mitochondria with PLA2G6 dysfunction. Acta Neuropathol. Commun. 2016, 4, 27. [Google Scholar] [CrossRef] [PubMed]
  148. Lin, G.; Lee, P.T.; Chen, K.; Mao, D.; Tan, K.L.; Zuo, Z.; Lin, W.W.; Wang, L.; Bellen, H.J. Phospholipase PLA2G6, a Parkinsonism-Associated Gene, Affects Vps26 and Vps35, Retromer Function, and Ceramide Levels, Similar to α-Synuclein Gain. Cell Metab. 2018, 28, 605–618. [Google Scholar] [CrossRef] [PubMed]
  149. Olsen, A.L.; Feany, M.B. Parkinson’s disease risk genes act in glia to control neuronal α-synuclein toxicity. Neurobiol. Dis. 2021, 159, 105482. [Google Scholar] [CrossRef]
  150. Hatton, C.; Ghanem, S.S.; Koss, D.J.; Abdi, I.Y.; Gibbons, E.; Guerreiro, R.; Bras, J.; International DLB Genetics Consortium; Walker, L.; Gelpi, E.; et al. Prion-like α-synuclein pathology in the brain of infants with Krabbe disease. Brain 2022, 145, 1257–1263. [Google Scholar] [CrossRef]
  151. Suzuki, K.; Iseki, E.; Togo, T.; Yamaguchi, A.; Katsuse, O.; Katsuyama, K.; Hirayasu, Y. Neuronal and glial accumulation of a- and b-synucleins in human lipidoses. Acta Neuropathol. 2007, 114, 481–489. [Google Scholar] [CrossRef]
  152. Hamano, K.; Hayashi, M.; Shioda, K.; Fukatsu, R.; Mizutani, S. Mechanisms of neurodegeneration in mucopolysaccharidoses II and IIIB: Analysis of human brain tissue. Acta Neuropathol. 2008, 115, 547–559. [Google Scholar] [CrossRef] [PubMed]
  153. Martinez, Z.; Zhu, M.; Han, S.; Fink, A.L. GM1 specifically interacts with alpha-synuclein and inhibits fibrillation. Biochemistry 2007, 46, 1868–1877. [Google Scholar] [CrossRef] [PubMed]
  154. Fantini, J.; Yahi, N. Molecular basis for the glycosphingolipid-binding specificity of a-synuclein: Key role of tyrosine 39 in membrane insertion. J. Mol. Biol. 2011, 408, 654–669. [Google Scholar] [CrossRef] [PubMed]
  155. Grey, M.; Dunning, C.J.; Gaspar, R.; Grey, C.; Brundin, P.; Sparr, E.; Linse, S. Acceleration of a-synuclein aggregation by exosomes. J. Biol. Chem. 2015, 290, 2969–2982. [Google Scholar] [CrossRef]
  156. De Franceschi, G.; Frare, E.; Bubacco, L.; Mammi, S.; Fontana, A.; Polverino de Laureto, P. Molecular insights into the interaction between α-synuclein and docosahexaenoic acid. J. Mol. Biol. 2009, 394, 94–107. [Google Scholar] [CrossRef]
  157. Sharon, R.; Goldberg, M.S.; Bar-Josef, I.; Betensky, R.A.; Shen, J.; Selkoe, D.J. α-Synuclein occurs in lipid-rich high molecular weight complexes, binds fatty acids, and shows homology to the fatty acid-binding proteins. Proc. Natl. Acad. Sci. USA 2001, 98, 9110–9115. [Google Scholar] [CrossRef]
  158. Lücke, C.; Gantz, D.L.; Klimtchuk, E.; Hamilton, J.A. Interactions between fatty acids and α-synuclein. J. Lipid Res. 2006, 47, 1714–1724. [Google Scholar] [CrossRef]
  159. Broersen, K.; van den Brink, D.; Fraser, G.; Goedert, M.; Davletov, B. α-Synuclein adopts an α-helical conformation in the presence of polyunsaturated fatty acids to hinder micelle formation. Biochemistry 2006, 45, 15610–15616. [Google Scholar] [CrossRef]
  160. De Franceschi, G.; Polverino de Laureto, P. Role of different regions of α-synuclein in the interaction with the brain fatty acid DHA. J. Chromatogr. Sep. Tech. 2014, 5, 219–226. [Google Scholar] [CrossRef]
  161. Sharon, R.; Bar-Joseph, I.; Frosch, M.P.; Walsh, D.M.; Hamilton, J.A.; Selkoe, D.J. The formation of highly soluble oligomers of α-synuclein is regulated by fatty acids and enhanced in Parkinson’s disease. Neuron 2003, 37, 583–595. [Google Scholar] [CrossRef]
  162. De Franceschi, G.; Frare, E.; Pivato, M.; Relini, A.; Penco, A.; Greggio, E.; Bubacco, L.; Fontana, A.; Polverino de Laureto, P. Structural and morphological characterization of aggregated species of α-synuclein induced by docosahexaenoic acid. J. Biol. Chem. 2011, 286, 22262–22274. [Google Scholar] [CrossRef] [PubMed]
  163. Necula, M.; Chirita, C.N.; Kuret, J. Rapid anionic micelle-mediated α-synuclein fibrillization in vitro. J. Biol. Chem. 2003, 278, 46674–46680. [Google Scholar] [CrossRef]
  164. Iljina, M.; Tosatto, L.; Choi, M.L.; Sang, J.C.; Ye, Y.; Hughes, C.D.; Bryant, C.E.; Gandhi, S.; Klenerman, D. Arachidonic acid mediates the formation of abundant α-helical multimers of α-synuclein. Sci. Rep. 2016, 6, 33928. [Google Scholar] [CrossRef]
  165. Perrin, R.J.; Woods, W.S.; Clayton, D.F.; George, J.M. Exposure to long chain polyunsaturated fatty acids triggers rapid multimerization of synucleins. J. Biol. Chem. 2001, 276, 41958–41962. [Google Scholar] [CrossRef] [PubMed]
  166. Riedel, M.; Goldbaum, O.; Wille, M.; Richter-Landsberg, C. Membrane lipid modification by docosahexaenoic acid (DHA) promotes the formation of α-synuclein inclusion bodies immunopositive for SUMO-1 in oligodendroglial cells after oxidative stress. J. Mol. Neurosci. 2011, 43, 290–302. [Google Scholar] [CrossRef] [PubMed]
  167. Almandoz-Gil, L.; Welander, H.; Ihse, E.; Khoonsari, P.E.; Musunuri, S.; Lendel, C.; Sigvardson, J.; Karlsson, M.; Ingelsson, M.; Kultima, K.; et al. Low molar excess of 4-oxo-2-nonenal and 4-hydroxy-2-nonenal promote oligomerization of alpha-synuclein through different pathways. Free Radic. Biol. Med. 2017, 110, 421–431. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of three regions of the αSyn protein. The N-terminal region binds to lipid membranes (blue). The non-amyloid-β component (NAC) region is the core of amyloid-like fibrils (red). The C-terminal region interacts with the N-terminal and NAC regions (green). The N-terminal region has tandem repeat motifs consisting of the KTKEGV sequence. Arrowheads indicate αSyn mutations responsible for fPD.
Figure 1. Schematic diagram of three regions of the αSyn protein. The N-terminal region binds to lipid membranes (blue). The non-amyloid-β component (NAC) region is the core of amyloid-like fibrils (red). The C-terminal region interacts with the N-terminal and NAC regions (green). The N-terminal region has tandem repeat motifs consisting of the KTKEGV sequence. Arrowheads indicate αSyn mutations responsible for fPD.
Ijms 25 08935 g001
Figure 2. Schematic model of the aggregation process of αSyn. Lipids interact with αSyn and possibly modulate its LLPS-induced liquid droplet formation, which may eventually trigger its aggregation.
Figure 2. Schematic model of the aggregation process of αSyn. Lipids interact with αSyn and possibly modulate its LLPS-induced liquid droplet formation, which may eventually trigger its aggregation.
Ijms 25 08935 g002
Table 1. Relationship between αSyn mutation, lipid-binding affinity, and aggregation.
Table 1. Relationship between αSyn mutation, lipid-binding affinity, and aggregation.
αSyn MutationsLipid-Binding AffinityAggregation
A30P
E46K
H50Q
G51D
A53T
A53E
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

Kamano, S.; Ozawa, D.; Ikenaka, K.; Nagai, Y. Role of Lipids in the Pathogenesis of Parkinson’s Disease. Int. J. Mol. Sci. 2024, 25, 8935. https://doi.org/10.3390/ijms25168935

AMA Style

Kamano S, Ozawa D, Ikenaka K, Nagai Y. Role of Lipids in the Pathogenesis of Parkinson’s Disease. International Journal of Molecular Sciences. 2024; 25(16):8935. https://doi.org/10.3390/ijms25168935

Chicago/Turabian Style

Kamano, Shumpei, Daisaku Ozawa, Kensuke Ikenaka, and Yoshitaka Nagai. 2024. "Role of Lipids in the Pathogenesis of Parkinson’s Disease" International Journal of Molecular Sciences 25, no. 16: 8935. https://doi.org/10.3390/ijms25168935

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