Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2010 Jul 1.
Published in final edited form as: Biochim Biophys Acta. 2008 Oct 9;1792(7):616–624. doi: 10.1016/j.bbadis.2008.09.013

Molecular Mechanisms of α-Synuclein Neurodegeneration

Elisa A Waxman 1, Benoit I Giasson 1
PMCID: PMC2756732  NIHMSID: NIHMS128251  PMID: 18955133

Abstract

α-Synuclein is an abundant highly charged protein that is normally predominantly localized around synaptic vesicles in presynatic terminals. Although the function of this protein is still ill-defined, genetic studies have demonstrated that point mutations or genetic alteration (duplications or triplications) that increase the number of copies of the α-synuclein (SCNA) gene can cause Parkinson’s disease or the related disorder dementia with Lewy bodies. α-Synuclein can aberrantly polymerize into fibrils with typical amyloid properties, and these fibrils are the major component of many types of pathological inclusions, including Lewy bodies, which are associated with neurodegenerative diseases, such as Parkinson’s disease. Genetic studies have clearly established that alteration in the α-synuclein gene can lead to neuronal demise. Although there is substantial evidence supporting the toxic nature of α-synuclein inclusions, other modes of toxicity such as oligomers have been proposed. In this review, some of the evidence for the different mechanisms of α-synuclein toxicity is presented and discussed.

Keywords: α-synuclein, amyloid, fibrils, Parkinson’s disease, protofibrils, toxicity

Introduction

Parkinson’s disease (PD) is a progressive, neurodegenerative disorder characterized by bradykinesia, resting tremor, cogwheel rigidity and postural instability [1,2] associated with the loss of dopaminergic (DA) neurons in the substantia nigra (SN) pars compacta [3,4]. Although the major pathological hallmarks of PD, Lewy bodies (LB) and Lewy neurites (LN), were originally observed in 1912 [5], α-synuclein (α-syn) was not identified as the major component of these proteinaceous inclusions until 1997 [6] following the discovery of PD kindred with point mutations in the α-syn gene (SNCA) [7]. In addition to PD, the presence of α-syn pathological inclusions is one of the defining features of several other neurodegenerative diseases, including dementia with Lewy bodies (DLB), LB variant of Alzheimer’s disease and multiple system atrophy (MSA) [6,813]. Furthermore α-syn inclusions are also found in a significant percentage of other neurodegenerative disorders, including neurodegeneration with brain iron accumulation type-1, Down’s syndrome, progressive autonomic failure and familial and sporadic Alzheimer’s disease [1421]. Collectively these diseases have been defined as α-synucleinopathies. This review will focus on the potential molecular mechanisms by which α-syn may cause neurodegeneration.

The α-Synuclein Protein

α-Syn is a small, highly charged 140-amino acid residue protein characterized by several major regions: 1) an amino-terminal region containing several imperfect KTKEGV repeats, 2) a hydrophobic center domain also referred to as the non-amyloid component (NAC) region, and 3) a highly negatively charged carboxy-terminus region (Figure 1). α-Syn is a soluble, heat-stable and natively “unfolded” protein [22,23]. It is predominantly expressed in central nervous system (CNS) neurons, where it is localized at presynaptic terminals in close proximity to synaptic vesicles [2427] and can associate with lipid membranes by forming amphiphatic α-helices, as shown in vitro [22,2831]. Although the function of α-syn is still poorly understood, several studies suggest that it is involved in modulating synaptic transmission, the density of synaptic vesicles and neuronal plasticity [26,27,3234], as well as provide a supportive role in the folding/refolding of SNARE proteins critical for neurotransmitter release, vesicle recycling and synaptic integrity [35]. However, knockout mouse models of α-syn are not lethal, and brain morphology is intact, suggesting that α-syn is not required for neuronal development and/ or that compensatory pathways are present [33,34]. In vitro studies have shown that the carboxy-terminal region of α-syn is required for chaperone-like activity [3638]. α-Syn can also associated with many proteins [39] and can regulate the activity of several enzymes, including tyrosine hydroxylase, the rate-limiting enzyme in dopamine production [40,41], mitogen-activated protein kinases (MAPKs) [42], and phospholipase D (PLD) [43,44].

Figure 1. Amino acid sequence and regions of α-synuclein.

Figure 1

α-Syn is composed of: 1) an amino-terminal domain (black) containing several imperfect KTKEGV motifs (blue underline); 2) a hydrophobic center (purple) termed non-amyloid component (NAC); and 3) a negatively charged carboxy-terminus (green). Three familial mutations in α-syn (red) have been identified in patients with PD.

Disease-Causing α-Synuclein Mutations

The most direct and compelling evidence for a fundamental role of α-syn in the pathogenesis of α-synucleionopathies is the causal relationship between genetic mutations and disease. The mutation Ala53Thr resulting from a G to A nucleotide transition at position 209 of the SNCA (α-syn) gene was first identified in a large Italian family (Contursi) and three small Greek families with autosomal dominant PD [7]. Thereafter, the Ala53Thr mutation was identified in at least 8 additional PD kindreds [4548], and another autosomal dominant mutation (Ala30Pro) was identified in a German kindred [49]. The Glu46Lys mutation in α-syn was identified in kindred manifesting classical PD or PD with features of dementia associated with widespread Lewy pathology, referred to as DLB [50]. In addition, short chromosomal duplications or trisomies containing the SCNA gene, plus relatively short flanking regions on chromosome 4, were discovered in patients with PD or DLB [5153], indicating that a 50 % increase in the expression of α-syn is sufficient to cause disease.

Formation of α-Synuclein Amyloidogenic Inclusions

α-Syn filaments (10–15 nm-wide) are the major ultrastructural component of pathological inclusions characteristic of synucleinopathies [6,12,13]. These inclusions can occur in cell bodies such as LBs that are present in neurons of patients with PD and DLB or such as glial cytoplasmic inclusions (GCIs) that form in oligodendrocytes of individuals with MSA. α-Syn inclusions can also present in the processes of affected cells, forming for example dystrophic neurites or large axonal swellings termed neuroaxonal spheroids [12,13].

In vitro studies have shown that recombinant soluble α-syn can readily polymerize into amyloidogenic fibrils that are structurally similar to those in human brains [5457]. Polymerization of α-syn is a nucleation dependent process, i.e. fibril formation displays a lag phase followed by a rapid increase in fibril formation [5760], and this lag phase is dramatically reduced by the addition of a “seed” or “nucleus” of pre-aggregated α-syn [59]. The polymerization of α-syn is associated with a dramatic conformational change from random coiled to predominantly β-pleated sheet [6164], and it has been proposed that α-syn progresses from an unordered monomer through partially folded intermediates, and finally elongates into “mature” filaments [58].

The central hydrophobic region in α-syn is necessary for fibrillization [64,65] and this region is buried within the fibril core, as demonstrated by immuno-electron microscopy analysis and proteinase K resistance assays [64,66,67]. The negatively charged carboxy-terminus negatively regulates fibril formation such that deletions of this region promotes fibril formation [62,67]. The presence of the amino-terminal region also reduces fibrillogenesis, as deletions of some of the repeats can accelerate filament formation [68].

The Ala53Thr and Glu46Lys α-syn proteins show increased rates of self-assembly and fibril formation [5455,57,60,6972]. This is consistent with studies showing that the Ala53Thr α-syn preferentially adopts a β-sheet conformation [73]. These in vitro data suggest that the Ala53Thr and Glu46Lys α-syn mutations could be pathogenic because of their increased the propensity to form pathological inclusions. In fact, α-syn, which contains both the Gly46Lys and Ala53Thr mutations, fibrillizes more rapidly than the Ala53Thr mutation alone, and this protein is even more likely to conform to the amino-terminal structure of α-syn in pathological inclusions, as detected by conformational-specific antibodies [74].

While some reports suggest that Ala30Pro α-syn forms fibrils more slowly than WT α-syn [69], this finding is not consistently observed by others [55,57]. In addition, the Ala30Pro mutation appears to affect α-syn properties independent of protein aggregation. The Ala30Pro mutation may partially impaired the ability of α-syn to bind to brain vesicles [35,75], likely due to a decreased likelihood to form α-helices [35,73]. However, it does not significantly prevent α-syn localization to presynaptic terminals [76]. This mutation can also directly impair the in vitro chaperone-like activity of α-syn [36], and studies in mice indicate that Ala30Pro α-syn is deficient in the ability of refolding SNARE proteins [35], which may be due to its reduced ability to interact with vesicles.

Protofibrils/ Oligomers Toxicity Hypothesis

The polymerization of α-syn from unstructured monomer to mature amyloid fibrils rich in β-sheets proceeds through the formation of several altered-sized oligomers and polymers that can be visualized and assayed by electron microscopy, atomic force microscopy and size-exclusions chromatography (Figure 2) [69,77]. Several of these intermediates (as well as products that may not culminant into fibrils) have been described as spheres (2–6 nm in size), chains of spheres (also termed protofibrils) and rings resembling circular protofibrils (also termed annular protofibrils) [69,77,78].

Figure 2. Formation of α-synuclein polymeric intermediates and fibrils.

Figure 2

α-Syn in native, monomeric form is mostly unstructured. Under certain conditions α-syn can undergo structural changes, resulting in β-pleated sheet formation. This form of α-syn can take two pathways, one which is off of the fibrillar pathway, and the other which will eventually form mature fibrils. The off-fibril pathway can result in the formation of annular or other forms of oligomers that will never develop into mature fibrils. The fibrillar pathway undergoes intermediate stages, which include protofibrils, before maturing into long strands and becoming LBs or LNs.

Several findings have suggested that protofibrils or some forms of α-syn oligomers may be toxic. This hypothesis is analogous to the proposal that amyloid assembly intermediates of other amyloidogenic proteins such as the Aβ peptide involved in Alzheimer’s disease may be toxic [77,79]. The initial observation that Ala30Pro α-syn may have a tendency to accumulate as oligomers instead of mature fibrils led to the suggestion that α-syn may have a similar toxic mechanism [69,77]. In addition, the formation of such oligomers in vitro is shown to increase leakiness of synthetic lipid vesicles [80]. Two mechanisms have been proposed to explain this effect of α-syn oligomers on membrane permeability: 1) α-syn annular oligomers may integrate into membrane resulting in the formation of pores or channel-like structures that could cause uncontrolled membrane permeability [7779,8183], and 2) oligomers enhance the ability of ions to move through the membrane bilayer, without the formation of pores [84].

Direct in vivo data supporting the “toxic oligomeric α-syn hypothesis” are still relatively limited and most of the evidence is circumstantial. There are studies in cultured cells that support this notion, but others demonstrated a lack of association between intracellular oligomer and toxicity (see the section on “Cell culture studies of α-synuclein toxicity” below). There is also a paucity of animal studies to directly support this hypothesis, but this could be related to the difficulties in monitoring these species in vivo due to their transitory nature.

Cell Culture Studies of α-Synuclein Toxicity

Several studies in cultured cells indicate that the expression of mutants of α-syn (Ala53Thr or Ala30Pro) can sensitize cells to toxic challenges; however, the results of these studies are not unequivocal [8596]. In some studies, toxicity induced by the expression of mutant α-syn is associated with the formation of α-syn aggregates [88,89,91,97]. The over-expression of wild-type (WT) α-syn in some settings has also been reported to render cells more vulnerable to cellular challenges [42,94,97103]. In striking contrast, other studies have shown that expression of WT α-syn can protect against cellular stresses [88,99,104109]. Nevertheless, several different mechanisms, including proteasomal inhibition, affects on signal transduction pathways, mitochondrial alterations, increased levels of free radicals, and membrane clustering of dopamine transporter resulting in increased dopamine uptake, have been proposed as toxic mechanisms associated with the expression of WT or mutant α-syn [42,90,92,96,98100,110].

Several studies have shown that the extracellular addition of in vitro generated α-syn oligomers to cultured cells can lead to toxicity [111,112]. Although these studies are artificial, there is some data to suggest that α-syn that may normally be secreted by cells or released due to cell death can be directly or indirectly toxic to other adjacent cells (see [113] and references therein). Some studies suggest that the intracellular formation of α-syn oligomers in H4 neuroglioma cells is associated with toxicity [114]. Conversely, others studies have shown that the formation of abundant Ala53Thr α-syn oligomers in SH-SY5Y cells induced by increasing intracellular catecholamine levels is not associated with toxicity [115].

The reasons for the discrepancies in the results between the toxic or protective effects of α-syn are not clear, but cell types, the promoters and transfection methods used to overexpress α-syn, the use of tagged protein versus native α-syn, the nature of the toxic stimulus utilized, and the level of expression, may be important factors [110].

Studies of α-Syn Toxicity in Yeast (Sacchatomyes Cerevisiae)

Although no orthologue of α-syn exist in yeast, expression of untagged WT α-syn or WT α-syn-EGFP in the yeast Sacchatomyes cerevisiae (S. cerevisiae) can inhibit cell growth and may result in cell death [116119]. When expressed at low levels or upon initial expression at high levels, α-syn is localized to the plasma membrane, but when expressed at high levels it subsequently forms cytoplasmic inclusions that is associated with toxicity [116118]. Expression of Ala53Thr α-syn results in similar distribution profiles and toxicity. In contrast, Ala30Pro α-syn displays both plasma membrane and diffuse cytoplasmic localizations, does not form inclusions, and demonstrates much reduced cell growth inhibition [116,117]. The altered distribution of Ala30Pro α-syn could be due to its reduced affinity for lipid membranes and/or targeting to the vacuole for degradation [119]. Further studies show that the α-syn inclusions in yeast are not comprised of amyloid-like fibrils, but instead are α-syn associated with clusters of vesicles [116,120]. The expression of α-syn in S. cerevisiae impairs endoplasmic reticulum (ER) to Golgi vesicle trafficking, which leads to the accumulation of these membranous vesicles [116,118,120]. A genome-wide screen identified several suppressors of these defects in ER-Golgi trafficking, including the Rab guanosine triphosphatase Ypt1p, which also suppresses α-syn toxicity [118]. Although the importance of these findings as it relates to the pathobiology of α-syn in humans is not completely clear, there is some experimental evidence suggesting that the deleterious effects of α-syn in this pathway may be relevant to disease. Expression of Rab1 (the murine othrolog of YPT1) in Drosophila rescues toxicity induced by expressing WT and Ala53Thr α-syn in DA neurons [118]. Similarly, expression of Rab1 can rescue the demise of DA neurons induced by over-expressing Ala53Thr α-syn in cultured rat midbrain primary neurons [118].

Studies of α-Syn Toxicity in Drosophila Melanogaster

There are no known orthologs of α-syn in Drosophila Melanogaster; nevertheless the ability to use this organism to identify genetic modifiers and to conduct studies of neurodegeneration in a shorter timeframe than in mammals has compelled the development of α-syn transgenic (tg) flies. Expression of human α-syn in Drosophila is reported to result in a selective age-dependent neuronal loss of DA neurons, locomotor dysfunction and cytoplasmic inclusions, some of which are composed to 7–10 nm wide filaments with additional granular material similar to LBs [121]. Co-expression of the chaperone heat-shock protein (Hsp) 70 can suppress neuronal degeneration, while a dominant negative form of Hsp 70 increases DA neuronal loss [122]. In addition, a drug that inhibits Hsp 90, a negative regulator of heat shock response, also rescues DA neuronal loss. However, these initial observations in α-syn tg flies have been challenged since locomotor dysfunction could not be replicated by others [123]. In addition, the loss of DA neurons is not observed when a whole-mount immunohistochemistry approach is used, compared to sequential paraffin sectioning used in the other studies [123], suggesting that expression of α-syn may not result in DA neuronal death but in other toxic effects that could alter neuronal morphology or the size of DA neurons.

Studies of α-Syn Toxicity in Caenorhabditis elegans

There is also no known ortholog of α-syn in the worm Caenorhabditis elegans (C. elegans), but the potential utility of this organism to quickly identify genetic modifiers compelled studies to develop α-syn tg models. However, the effects of expressing WT and mutants of α-syn in DA neurons of C. elegans have been controversial. Lakso and colleagues report that expression of WT or A53T human α-syn using pan-neuronal or DA neuronal promoters cause a reduction in the number of DA neuron cell bodies and processes [124]. Conversely, Kuwahara and colleagues do not observe a demise of DA neurons using a DA-specific promoter to express human WT, A30P or A53T α-syn, although a decrease of neurites is noted [125]. It is suggested that the apparent reduction in DA neurons observed by Lakso and colleagues could be due to a reduction in the expression of the tg-expressed marker used to monitor DA neurons, resulting from using the same promoter to express α-syn [125]. Nevertheless, Kowahara and colleagues report that the tg expression of A53T or A30P α-syn results in a reduction in DA levels associated with impairment in locomotor rate in response to food, which in C. elegans is attributed to the function of DA neurons [125].

Mouse Tg Models of α-Syn Toxicity

Several tg mouse models expressing either WT or mutant (Ala53Thr and Ala30Pro) α-syn have been reported. Masliah and colleagues reported on the first α-syn tg mouse lines that were generated [126]. In these mice neuronal expression of WT human α-syn is driven using a platelet-derived growth factor-β (PDGF-β) promoter, which results in the formation of amorphous, non-filamentous α-syn neuron aggregates in the neocortex, the hippocampus, and occasionally in the SN. A subset of α-syn inclusions is also ubiquitin positive, which is characteristic of authentic human α-syn inclusions. However, in contrast with typical α-syn inclusions in PD, a significant portion of the inclusions in these mice are located in the nucleus. The formation α-syn aggregates in PDGF-β/α-syn tg mice is also associated with a modest reduction in striatal tyrosine hydroxylase-positive terminals. Interestingly, over-expression of β-syn, a protein with close homology to α-syn, but unable to polymerize into amyloid fibrils [64], by transgenic cross breeding, reduces the numbers of α-syn inclusions, ameliorates motor impairment and results in a partial rescue of striatal tyrosine hydroxylase-positive terminals suggesting that β-syn may prevent α-syn aggregation [127].

Since Alzheimer’s disease pathology and PD often coincide in patients, the effects of accumulating amyloid-β (Aβ) peptide, the major component of senile plaques characteristic of Alzheimer’s disease, on α-syn pathobiology has been assessed using tg mice. The PDGF-β/α-syn tg mice have been cross-bred with a tg mouse model of Alzheimer’s disease line, where a disease-causing mutant form of the human amyloid precursor protein (APP) is expressed resulting in the production and accumulation Aβ extracellular inclusions [128]. The expression of Aβ peptide in these bigenic mice is shown to potentiate neuronal and presynaptic terminal loss, motor impairments and the formation of fibrillar intraneuronal α-syn when compared to PDGF-β/α-syn tg mice. These findings provide important information supporting the notion that Aβ peptide, which is predominantly extracellular, can promote the formation of intraneuronal α-syn aggregates.

The PDGF-β/α-syn tg mice also have been used to generate bigenic tg mice that overexpress rat Hsp 70, and Hsp 70 expression is found to mitigate the formation of α-syn aggregates, suggesting that Hsp 70 may have a role in refolding or degrading misfolded α-syn [129].

Other studies have used a Thy-1 promoter to drive the neuronal expression of WT, Ala30Pro or Ala53Thr human α-syn in tg mice [76,130132]. In some of the Thy-1 tg mouse lines, expression of WT or Ala53Thr α-syn results in the appearance of perikaryal and neuritic accumulations of α-syn and age dependent motor impairment associated with the degeneration of ventral root axons and muscle denervation [130]. A subset of α-syn inclusions in these mice are argyrophilic and immunoreactive for ubiquitin, but they lack the filamentous characteristics of authentic human α-syn inclusions. Kahle and colleagues generated Thy-1 tg mouse lines expressing human WT or Ala30Pro α-syn. These Thy-1/WT-α-syn and Thy-1/Ala30Pro-α-syn tg mice are reported to initially developed detergent-insoluble somatodendritic accumulations of human α-syn that is not associated with any obvious phenotype [76,131,133]. With aging the Thy-1/Ala30Pro-α-syn tg mice develop neuronal cytoplasmic fibrillar “amyloidogenic” inclusions that resemble the properties of authentic α-syn inclusions as evidenced by thioflavin S-reactivity, proteinase K resistance and ultrastructure studies [132]. These inclusions are predominantly abundant in the midbrain, brainstem, amygdala and spinal cord [132,133]. From these studies reported by Kahle and colleagues, it is unclear if the Thy-1/WT-α-syn tg mice develop age-dependent pathological features similar to the Thy-1/Ala30Pro-α-syn transgenic mice or if these changes are specific and exacerbated by the Ala30Pro mutation [131,132]. The formation of amyloidogenic inclusions in Thy-1/Ala30Pro-α-syn tg mice is associated with deterioration in locomotor performance that progressed to spastic paralysis of the hind limbs [128]. Furthermore, the specific formation of amyloidogenic inclusions in the amygdale of Thy-1/Ala30Pro-α-syn is correlated with a decline in cognitive performance [133].

Tg mice have also been generated that expressed either WT, Ala53Thr or Ala30Pro human α-syn using the murine prion protein promoter (PrP) [134,135]. PrP α-syn tg mice expressing Ala53Thr α-syn, but not those expressing equivalent levels of WT or Ala30Pro α-syn, develop amyloidogenic, 10–15 nm filamentous α-syn inclusions in neurons (i.e. axonal spheroids, LB-like and LN-like lesions) that replicate many of the biochemical and histological features of authentic human synucleinopathies [134,135]. These α-syn inclusions are predominantly observed in the spinal cord, brain stem, deep cerebellar nuclei, and the thalamus. Similar to Thy-1/Ala30Pro α-syn transgenic mice, the hippocampus and the SN are spared. Also similar to Thy-1/Ala30Pro α-syn transgenic mice, PrP/Ala53Thr α-syn transgenic mice display an age-dependent severe motor phenotype that includes reduced ambulance, paralysis of the extremities usually beginning at a hind limb that progress to quadriparesis and arched back posture. These phenotypic features coincide with the accumulation of filamentous intracytoplasmic α-syn neuronal inclusions. The degeneration of motor axons is likely the main cause of the motor phenotype in the PrP/Ala53Thr a-syn tg mice, as dramatic Wallerian degeneration of ventral roots was observed [134]. Ultrastructure studies show that α-syn filamentous axonal inclusions can trap mitochondria and impair axonal transport leading to axonal swelling containing vacuoles, vesicles and mitochondria [134]. The formation of α-syn inclusions is also associated with motor neuron loss [136]. The increase propensity of Ala53Thr α-syn to polymerize into fibrils compared to WT or Ala30Pro α-syn [54,55,57] is the most likely explanation for the formation of neurotoxic inclusions in PrP/Ala53Thr α-syn tg mice, while PrP/WT α-syn tg or PrP/Ala30Pro α-syn tg mice with similar expression do not display pathology. Since the age-dependent phenotypic changes and pathologies in Thy/Ala30Pro α-syn tg mice are similar to those in PrP/Ala53Thr α-syn tg mice, but PrP/Ala30Pro α-syn tg mice are not affected, it is possible that Thy/Ala30Pro α-syn tg mice have higher expression levels than PrP/Ala30Pro α-syn tg mice. However, a direct comparison has not been performed. This possibility is further supported by PrP/Ala30Pro α-syn tg mice by Sudhöf and colleagues that display similar pathological and phenotypic features as described above for PrP/Ala53Thr α-syn tg mice [35,137].

Using the PrP/Ala30Pro α-syn tg mice, a dramatic increase in the level of the lipid-binding protein Apolipoprotein E (ApoE) coincides with the motor impairment and motor neuron loss associated with α-syn inclusions [137]. These findings are particularly interesting since specific ApoE genotypic isoforms are important risk factors for Alzheimer’s disease. Further, ApoE can modulate the formation of α-syn pathology, since bleeding these PrP/Ala30Pro α-syn tg mice on an ApoE null background delays motor disease, while increasing survival, although these processes are not completely abolished [137]. These findings indicate that ApoE can be an important modulator of α-syn aggregation and related pathogenesis, although the mechanisms are not clearly defined.

Since septin-4 (Sept4), a member of the septins family of polymerizing GTP binding proteins that serve as scaffolds that can anchor or stabilize other molecules, is present in α-syn pathological inclusions in human brains and interacts with α-syn [138], the effect of Sept4 on α-syn pathobiology has been investigated in vivo in the PrP/Ala53Thr α-syn tg mice described above. The breeding of PrP/Ala53Thr α-syn tg mice on a Sept4 null background results in an exacerbation of locomotor deterioration and neuronal loss associated with α-syn inclusion formation [139]. These data suggest that Sept4 may act as a suppressor of α-syn aggregation and resulting neurodegeneration.

Counter intuitively, the transgenic mice described above do not develop substantial SN DA neurons pathologies. For reasons that remain enigmatic, it appears that, in contrast to humans, this population of neurons in mice is resilient to the formation of α-syn inclusions and degeneration. This notion is consistent with the lack of pathology even when the a tyrosine hydroxylase promoter, which drives express specifically in catecholaminergic neurons, is used to generate WT, Ala53Thr or Ala30Pro α-syn tg mice [140].

Since α-syn inclusions in human brain contain C-terminally truncated α-syn [112,141143], which may be generated by incomplete 20 S proteasome degradation or calpain cleavage [112,144145], and C-terminal truncated α-syn fibrillizes faster in vitro (see above), Tofaris and colleagues created tg mice expressing C-terminal truncated α-syn [146]. These tg mice express human α-syn missing the last 20 amino acids (i.e. α-syn 1–120) driven by a rat tyrosine hydroxylase promoter on an α-syn null background. In this model, α-syn 1–120 is expressed in DA neurons of the SN and olfactory bulb, resulting in aggregates with either granular and fibrillar morphologies. Shrunken neuronal perikarya and swollen axons of DA neurons are observed, but without significant neuronal loss. In another effort to study the effect of truncated α-syn, transgenic mice were generated using rat tyrosine hydroxylase promoter to express Ala53Thr human α-syn 1–130 (i.e. lacking the last 10 amino acids) [147]. Animals that express this truncated protein have a significant loss of nigral DA neurons, which is not observed in a similar tg line expressing full-length human α-syn. However, the loss of DA neurons is shown to occur during embryogenesis without the formation of α-syn inclusion. These findings support that this form of truncated α-syn can be toxic in nature, but its does not provide insights in the typical age-dependent neurodegeneration associated with human diseases.

As GCIs in oligodendrocytes are key pathological features of MSA, several tg mouse lines expressing WT human α-syn in oligodendrocytes were generated. Kahle and colleagues used a proteolipid protein (PLP) promoter to drive express α-syn specifically in oligodendrocytes, and these mice develop detergent-insoluble aggregates with histological profiles that resembled GCIs [148]. These mice demonstrate increased sensitivity to striatonigral degeneration and olivopontocerebellar atrophy induced by the mitochondrial inhibitor 3-nitropropionic acid [149]. Expression of human α-syn using a murine myelin basic protein (MBP) promoter results in the specific expression of α-syn in oligodendroctyes with progressive accumulation of filamentous inclusions associated with disruption of myelin lamina and demyelination [150]. In these mice, the accumulation of α-syn in oligodendrocytes results in decreased dendritic density and loss of DA projections in the basal ganglia. In one of the MBP/α-syn tg mouse lines expressing the highest levels of α-syn, severe neurological features, including ataxia and seizure activity, resulting in premature death, is observed.

Transgenic mice expressing WT human α-syn in oligodendrocytes also have been generated using a murine 2’, 3’-cyclic nucleotide 3’-phosphodiesterase (CNP) promoter to drive expression [151]. The transgene is specifically expressed in oligodendroctyes resulting in age-dependent cytoplasmic brain and spinal cord accumulations and the formation of fibrillar inclusions. These inclusions are associated with demyelination, demise of oligodengrocytes, age-dependent motor impairment and brain atrophy. Injury to oligodendrocytes results in secondary neuronal degeneration including accumulation of perikaryal hyperphosphorylated neurofilaments, degeneration of axonal terminal, neuronal loss, and formation of neuronal inclusions comprised of endogenous mouse α-syn.

Recently, the first tetracyline-regulated conditional tg mice (tet-off) expressing human wt α-syn was described [152]. To drive inducible expression in specific neuronal populations these mice have been cross-bred to tg mice expressing tetracycline-controlled transactivator under the control of the hamster PrP or calcium/calmodulin-dependent protein kinase IIα (CaMKIIα) promoter. In one of these inducible PrP/α-syn tg mouse lines, α-syn is expressed in the olfactory bulb, cortex, basal ganglia and cerebellum, while in another inducible PrP/α-syn tg line expression is predominantly confined to olfactory bulb. In one of the inducible CaMKIIα/α-syn tg mouse lines, α-syn is expressed in the olfactory bulb, cortex, basal ganglia, hippocampus, thalamus and substantia nigra, including DA neurons. These CaMKIIα/α-syn tg mice exhibit reductions in SN DA neurons and hippocampal neurogenesis, without the presence of α-syn aggregates. They demonstrate a progressive motor decline as assayed by rotarod that can not be reversed by turning off tg α-syn expression.

Studies α-Syn Over-expression in Adult DA Neurons using Viral Delivery

Several studied have used viral vectors to express α-syn in adult rats or monkeys nigral DA neurons. One advantage of this approach is that it mitigates the possible effects of early developmental expression of α-syn with the possibility of compensatory mechanisms. Expression of human WT, Ala30Pro or Ala53Thr α-syn in rat or monkey nigral DA neurons using adeno-associated viral vectors that stably express the transgene (> 6 months) results in substantial and specific demise of these DA neurons (30–80% loss), concurrent with the formation of cellular α-syn inclusions and dystrophic neurites [153155]. Similar results are observed when using a lentiviral-based vector system to express α-syn proteins in rat nigral DA neurons; however, over-expression of rat α-syn is much less toxic [156].

General remarks

Although genetic and pathological studies have clearly demonstrated the importance of α-syn in the etiology of PD, several different mechanisms of toxicity have been proposed. These can be grouped into 3 categories based on the monomeric/polymeric nature of the proposed toxic species.

First, simple increases in intracellular abundance of monomeric α-syn have been proposed as a mode of neuronal toxicity. Some studies in cultured cells many support this notion; however, these findings are not unanimous and the loss of DA neurons in the SN or other types of neurons has not been observed in tg mice that simply over-express high levels of α-syn. In PD or DLB patients with duplication or triplication of the α-syn gene, where α-syn expression is increased by 50% or 100%, respectively, α-syn pathological inclusions always coincide with disease (see [157] and references therein). In addition, some studies have suggested that α-syn expression may be increased in specific brain areas or types of neurons in individuals with sporadic PD, but these findings have been challenged in other reports (see [158] and references therein).

Secondly, based on in vitro data discussed above, some forms of α-syn oligomers and protofibrils have been proposed as potent toxic specifies. However, this hypothesis still lacks solid direct in vivo studies documenting toxicity linked to the present of α-syn oligomers, although some models using cultured cell supports this notion. Conversely, biochemical studies have shown that the presence of some forms of α-syn oligomers in the midbrain of PrP/Ala53Thr α-syn tg mice without any evidence of toxicity to DA neurons [115].

Lastly, the notion that the aberrant polymerization of α-syn into filaments, which eventually form large intracytoplasmic inclusions, can cause the dysfunction and the demise of neurons or oligodendrocytes has been support by various experimental models, as described above. Furthermore, the involvement of α-syn aggregates in the dysfunction and demise of neurons is suggested by the correlations between severity of dementia and LB density in patients with DLB [159162]. In addition, it is likely that a profusion of smaller α-syn aggregates in the form of neuritic and pre-synaptic α-syn inclusions have a predominant role in impairing normal neuronal function [74,128,163165]. α-Syn aggregates may impair proteasome function [166], and they may act as "sinks," incidentally recruiting other necessary, cellular proteins from their normal cellular functions. α-Syn inclusions can impair cellular functions by obstructing normal cellular trafficking (including disruption of ER and Golgi apparatus), by disrupting cell morphology, by impairing axonal transport, and by trapping cellular components (eg. mitochondia).

It is important to emphasize that the different alternative mechanisms of α-syn toxicity based on the different forms of α-syn polymers are not necessarily mutually exclusive. The presence of any form of α-syn polymer, from small oligomers to amyloid fibrils, are abnormal and may be problematic for the normal activities of cells, thereby resulting in neurodegeneration.

ACKNOWLEDGEMENTS

This work was supported by grants from the National Institute on Aging (AG09215) and the National Institute of Neurological Disorders and Stroke (NS053488). E.A.W. was supported by a training grant (T32 AG00255) from the National Institute on Aging.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • 1.Simuni T, Hurtig HI. Parkinson's Disease: the Clinical Picture. In: Clark CM, Trojanoswki JQ, editors. Neurodegenerative dementias. New York: McGraw-Hill; 2000. pp. 193–203. [Google Scholar]
  • 2.Gelb DJ, Oliver E, Gilman S. Diagnostic criteria for Parkinson disease. Arch. Neurol. 1999;56:33–39. doi: 10.1001/archneur.56.1.33. [DOI] [PubMed] [Google Scholar]
  • 3.Pakkenberg B, Moller A, Gundersen HJ, Mouritzen DA, Pakkenberg H. The absolute number of nerve cells in substantia nigra in normal subjects and in patients with Parkinson's disease estimated with an unbiased stereological method. J. Neurol. Neurosurg. Psychiatry. 1991;54:30–33. doi: 10.1136/jnnp.54.1.30. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Damier P, Hirsch EC, Agid Y, Graybiel AM. The substantia nigra of the human brain. II. Patterns of loss of dopamine-containing neurons in Parkinson's disease. Brain. 1999;122(Pt 8):1437–1448. doi: 10.1093/brain/122.8.1437. [DOI] [PubMed] [Google Scholar]
  • 5.Holdorff B. Fritz Heinrich Lewy (1885–1950) J. Neurol. 253;2006:677–678. doi: 10.1007/s00415-006-0130-2. [DOI] [PubMed] [Google Scholar]
  • 6.Spillantini MG, Schmidt ML, Lee VMY, Trojanowski JQ, Jakes R, Goedert M. Alpha-synuclein in Lewy bodies. Nature. 1997;388:839–840. doi: 10.1038/42166. [DOI] [PubMed] [Google Scholar]
  • 7.Polymeropoulos MH, Lavedan C, Leroy E, Ide SE, Dehejia A, Dutra A, Pike B, Root H, Rubenstein J, Boyer R, Stenroos ES, Chandrasekharappa S, Athanassiadou A, Papapetropoulos T, Johnson WG, Lazzarini AM, Duvoisin RC, Di Iorio G, Golbe LI, Nussbaum RL. Mutation in the alpha-synuclein gene identified in families with Parkinson's disease. Science. 1997;276:2045–2047. doi: 10.1126/science.276.5321.2045. [DOI] [PubMed] [Google Scholar]
  • 8.Duda JE, Giasson BI, Gur TL, Montine TJ, Robertson D, Biaggioni I, Hurtig HI, Stern MB, Gollomp SM, Grossman M, Lee VMY, Trojanowski JQ. Immunohistochemical and biochemical studies demonstrate a distinct profile of alpha-synuclein permutations in multiple system atrophy. J. Neuropathol. Exp. Neurol. 2000;59:830–841. doi: 10.1093/jnen/59.9.830. [DOI] [PubMed] [Google Scholar]
  • 9.Lippa CF, Schmidt ML, Lee VM, Trojanowski JQ. Dementia with Lewy bodies. Neurology. 1999;52:893. doi: 10.1212/wnl.52.4.893. [DOI] [PubMed] [Google Scholar]
  • 10.Spillantini MG, Crowther RA, Jakes R, Cairns NJ, Lantos PL, Goedert M. Filamentous alpha-synuclein inclusions link multiple system atrophy with Parkinson's disease and dementia with Lewy bodies. Neurosci. Lett. 1998;251:205–208. doi: 10.1016/s0304-3940(98)00504-7. [DOI] [PubMed] [Google Scholar]
  • 11.Tu PH, Galvin JE, Baba M, Giasson B, Tomita T, Leight S, Nakajo S, Iwatsubo T, Trojanowski JQ, Lee VMY. Glial cytoplasmic inclusions in white matter oligodendrocytes of multiple system atrophy brains contain insoluble alpha-synuclein. Ann. Neurol. 1998;44:415–422. doi: 10.1002/ana.410440324. [DOI] [PubMed] [Google Scholar]
  • 12.Goedert M. Alpha-synuclein and neurodegenerative diseases. Nat. Rev. Neurosci. 2001;2:492–501. doi: 10.1038/35081564. [DOI] [PubMed] [Google Scholar]
  • 13.Forman MS, Lee VM, Trojanowski JQ. Nosology of Parkinson's disease: looking for the way out of a quackmire. Neuron. 2005;47:479–482. doi: 10.1016/j.neuron.2005.07.021. [DOI] [PubMed] [Google Scholar]
  • 14.Arai K, Kato N, Kashiwado K, Hattori T. Pure autonomic failure in association with human alpha-synucleinopathy. Neurosci. Lett. 2000;296:171–173. doi: 10.1016/s0304-3940(00)01623-2. [DOI] [PubMed] [Google Scholar]
  • 15.Arawaka S, Saito Y, Murayama S, Mori H. Lewy body in neurodegeneration with brain iron accumulation type 1 is immunoreactive for alpha-synuclein. Neurology. 1998;51:887–889. doi: 10.1212/wnl.51.3.887. [DOI] [PubMed] [Google Scholar]
  • 16.Galvin JE, Giasson B, Hurtig HI, Lee VMY, Trojanowski JQ. Neurodegeneration with brain iron accumulation, type 1 is characterized by alpha-, beta-, and gamma-synuclein neuropathology. Am. J. Pathol. 2000;157:361–368. doi: 10.1016/s0002-9440(10)64548-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Hamilton RL. Lewy bodies in Alzheimer's disease: a neuropathological review of 145 cases using alpha-synuclein immunohistochemistry. Brain Pathol. 2000;10:378–384. doi: 10.1111/j.1750-3639.2000.tb00269.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Kaufmann H, Hague K, Perl D. Accumulation of alpha-synuclein in autonomic nerves in pure autonomic failure. Neurology. 2001;56:980–981. doi: 10.1212/wnl.56.7.980. [DOI] [PubMed] [Google Scholar]
  • 19.Lippa CF, Fujiwara H, Mann DM, Giasson B, Baba M, Schmidt ML, Nee LE, O'Connell B, Pollen DA, George-Hyslop P, Ghetti B, Nochlin D, Bird TD, Cairns NJ, Lee VMY, Iwatsubo T, Trojanowski JQ. Lewy bodies contain altered alpha-synuclein in brains of many familial Alzheimer's disease patients with mutations in presenilin and amyloid precursor protein genes. Am. J. Pathol. 1998;153:1365–1370. doi: 10.1016/s0002-9440(10)65722-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Lippa CF, Schmidt ML, Lee VM, Trojanowski JQ. Antibodies to alpha-synuclein detect Lewy bodies in many Down's syndrome brains with Alzheimer's disease. Ann. Neurol. 1999;45:353–357. doi: 10.1002/1531-8249(199903)45:3<353::aid-ana11>3.0.co;2-4. [DOI] [PubMed] [Google Scholar]
  • 21.Wakabayashi K, Yoshimoto M, Fukushima T, Koide R, Horikawa Y, Morita T, Takahashi H. Widespread occurrence of alpha-synuclein/NACP-immunoreactive neuronal inclusions in juvenile and adult-onset Hallervorden-Spatz disease with Lewy bodies. Neuropathol. Appl. Neurobiol. 1999;25:363–368. doi: 10.1046/j.1365-2990.1999.00193.x. [DOI] [PubMed] [Google Scholar]
  • 22.Davidson WS, Jonas A, Clayton DF, George JM. Stabilization of alpha-synuclein secondary structure upon binding to synthetic membranes. J. Biol. Chem. 1998;273:9443–9449. doi: 10.1074/jbc.273.16.9443. [DOI] [PubMed] [Google Scholar]
  • 23.Weinreb PH, Zhen W, Poon AW, Conway KA, Lansbury PT. NACP, a protein implicated in Alzheimer's disease and learning, is natively unfolded. Biochemistry. 1996;35:13709–13715. doi: 10.1021/bi961799n. [DOI] [PubMed] [Google Scholar]
  • 24.George JM, Jin WS, Woods WS, Clayton DF. Characterization of a novel protein regulated during the critical period for song learning in the zebra finch. Neuron. 1995;15:361–372. doi: 10.1016/0896-6273(95)90040-3. [DOI] [PubMed] [Google Scholar]
  • 25.Jakes R, Spillantini MG, Goedert M. Identification of two distinct synucleins from human brain. FEBS Lett. 1994;345:27–32. doi: 10.1016/0014-5793(94)00395-5. [DOI] [PubMed] [Google Scholar]
  • 26.Withers GS, George JM, Banker GA, Clayton DF. Delayed localization of synelfin (synuclein, NACP) to presynaptic terminals in cultured rat hippocampal neurons. Brain Res. Dev. Brain Res. 1997;99:87–94. doi: 10.1016/s0165-3806(96)00210-6. [DOI] [PubMed] [Google Scholar]
  • 27.Iwai A, Masliah E, Yoshimoto M, Ge N, Flanagan L, de Silva HA, 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. doi: 10.1016/0896-6273(95)90302-x. [DOI] [PubMed] [Google Scholar]
  • 28.Clayton DF, George JM. Synucleins in synaptic plasticity and neurodegenerative disorders. J. Neurosci. Res. 1999;58:120–129. [PubMed] [Google Scholar]
  • 29.Mihajlovic M, Lazaridis T. Membrane-bound structure and energetics of alpha-synuclein. Proteins. 2008;70:761–778. doi: 10.1002/prot.21558. [DOI] [PubMed] [Google Scholar]
  • 30.Kim YS, Laurine E, Woods W, Lee SJ. A novel mechanism of interaction between alpha-synuclein and biological membranes. J. Mol. Biol. 2006;360:386–397. doi: 10.1016/j.jmb.2006.05.004. [DOI] [PubMed] [Google Scholar]
  • 31.Zhu M, Fink AL. Lipid binding inhibits alpha-synuclein fibril formation. J. Biol. Chem. 2003;278:16873–16877. doi: 10.1074/jbc.M210136200. [DOI] [PubMed] [Google Scholar]
  • 32.Murphy DD, Rueter SM, Trojanowski JQ, Lee VMY. Synucleins are developmentally expressed, and alpha-synuclein regulates the size of the presynaptic vesicular pool in primary hippocampal neurons. J. Neurosci. 2000;20:3214–3220. doi: 10.1523/JNEUROSCI.20-09-03214.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Cabin DE, Shimazu K, Murphy D, Cole NB, Gottschalk W, Mcllwain KL, Orrison B, Chen A, Ellis CE, Paylor R, Lu B, Nussbaum RL. Synaptic vesicle depletion correlates with attenuated synaptic responses to prolonged repetitive stimulation in mice lacking α-synuclein. J. Neurosci. 2002;22:8797–8807. doi: 10.1523/JNEUROSCI.22-20-08797.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Abeliovich A, Schmitz Y, Farinas I, Choi-Lundberg D, Ho WH, Castillo PE, Shinsky N, Verdugo JM, Armanini M, Ryan A, Hynes M, Phillips H, Sulzer D, Rosenthal A. Mice lacking alpha-synuclein display functional deficits in the nigrostriatal dopamine system. Neuron. 2000;25:239–252. doi: 10.1016/s0896-6273(00)80886-7. [DOI] [PubMed] [Google Scholar]
  • 35.Chandra S, Gallardo G, Fernández-Chacón R, Schlüter OM, Südhof TC. Alpha-synuclein cooperates with CSPalpha in preventing neurodegeneration. Cell. 2005;123:359–361. doi: 10.1016/j.cell.2005.09.028. [DOI] [PubMed] [Google Scholar]
  • 36.Souza JM, Giasson BI, Lee VM, Ischiropoulos H. Chaperone-like activity of synucleins. FEBS Lett. 2000;474:116–119. doi: 10.1016/s0014-5793(00)01563-5. [DOI] [PubMed] [Google Scholar]
  • 37.Kim TD, Paik SR, Yang CH, Kim J. Structural changes in alpha-synuclein affect its chaperone-like activity in vitro. Protein Sci. 2000;9:2489–2496. doi: 10.1110/ps.9.12.2489. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Kim TD, Paik SR, Yang CH. Structural and functional implications of C-terminal regions of alpha-synuclein. Biochemistry. 2002;41:13782–13790. doi: 10.1021/bi026284c. [DOI] [PubMed] [Google Scholar]
  • 39.Norris EH, Giasson BI, Lee VM. Alpha-synuclein: normal function and role in neurodegenerative diseases. Curr. Top. Dev. Biol. 2004;60:17–54. doi: 10.1016/S0070-2153(04)60002-0. [DOI] [PubMed] [Google Scholar]
  • 40.Perez RG, Waymire JC, Lin E, Liu JJ, Guo F, Zigmond MJ. A Role for alpha - Synuclein in the Regulation of Dopamine Biosynthesis. J. Neurosci. 2002;22:3090–3099. doi: 10.1523/JNEUROSCI.22-08-03090.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Peng XM, Tehranian R, Dietrich P, Stefanis L, Perez RG. α-Synuclein activation of protein phosphatase 2A reduces tyrosine hydroxylase phosphorylation in dopaminergic cells. J Cell Sci. 2005;118:3523–3530. doi: 10.1242/jcs.02481. [DOI] [PubMed] [Google Scholar]
  • 42.Iwata A, Maruyama M, Kanazawa I, Nukina N. alpha-Synuclein affects the MAPK pathway and accelerates cell death. J. Biol. Chem. 2001;276:45320–45329. doi: 10.1074/jbc.M103736200. [DOI] [PubMed] [Google Scholar]
  • 43.Jenco JM, Rawlingson A, Daniels B, Morris AJ. Regulation of phospholipase D2: selective inhibition of mammalian phospholipase D isoenzymes by a alpha- and beta-synuclein. Biochemistry. 1998;37:4901–4909. doi: 10.1021/bi972776r. [DOI] [PubMed] [Google Scholar]
  • 44.Ahn BH, Rhim H, Kim SY, Sung YM, Lee MY, Choi JY, Wolozin B, Chang JS, Lee YH, Kwon TK, Chung KC, Yoon SH, Hahn SJ, Kim MS, Jo YH, Min DS. alpha -Synuclein interacts with phospholipase D isozymes and inhibits pervanadate-induced phospholipase D activation in human embryonic kidney-293 Cells. J. Biol. Chem. 2002;277:12334–12342. doi: 10.1074/jbc.M110414200. [DOI] [PubMed] [Google Scholar]
  • 45.Athanassiadou A, Voutsinas G, Psiouri L, Leroy E, Polymeropoulos MH, Ilias A, Maniatis GM, Papapetropoulos T. Genetic analysis of families with Parkinson disease that carry the Ala53Thr mutation in the gene encoding alpha-synuclein. Am. J. Hum. Genet. 1999;65:555–558. doi: 10.1086/302486. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Markopoulou K, Wszolek ZK, Pfeiffer RF, Chase BA. Reduced expression of the G209A alpha-synuclein allele in familial Parkinsonism. Ann. Neurol. 1999;46:374–381. doi: 10.1002/1531-8249(199909)46:3<374::aid-ana13>3.0.co;2-9. [DOI] [PubMed] [Google Scholar]
  • 47.Papadimitriou A, Veletza V, Hadjigeorgiou GM, Patrikiou A, Hirano M, Anastasopoulos I. Mutated alpha-synuclein gene in two Greek kindreds with familial PD: incomplete penetrance? Neurology. 1999;52:651–654. doi: 10.1212/wnl.52.3.651. [DOI] [PubMed] [Google Scholar]
  • 48.Spira PJ, Sharpe DM, Halliday G, Cavanagh J, Nicholson GA. Clinical and pathological features of a Parkinsonian syndrome in a family with an Ala53Thr alpha-synuclein mutation. Ann. Neurol. 2001;49:313–319. [PubMed] [Google Scholar]
  • 49.Kruger R, Kuhn W, Muller T, Woitalla D, Graeber M, Kosel S, Przuntek H, Epplen JT, Schols L, Riess O. Ala30Pro mutation in the gene encoding alpha-synuclein in Parkinson's disease. Nat. Genet. 1998;18:106–108. doi: 10.1038/ng0298-106. [DOI] [PubMed] [Google Scholar]
  • 50.Zarranz JJ, Alegre J, Gomez-Esteban JC, Lezcano E, Ros R, Ampuero I, Vidal L, Hoenicka J, Rodriguez O, Atares B, Llorens V, Gomez TE, del Ser T, Munoz DG, de Yebenes JG. The new mutation, E46K, of alpha-synuclein causes Parkinson and Lewy body dementia. Ann. Neurol. 2004;55:164–173. doi: 10.1002/ana.10795. [DOI] [PubMed] [Google Scholar]
  • 51.Chartier-Harlin MC, Kachergus J, Roumier C, Mouroux V, Douay X, Lincoln S, Levecque C, Larvor L, Andrieux J, Hulihan M, Waucquier N, Defebvre L, Amouyel P, Farrer M, Destee A. Alpha-synuclein locus duplication as a cause of familial Parkinson's disease. Lancet. 2004;364:1167–1169. doi: 10.1016/S0140-6736(04)17103-1. [DOI] [PubMed] [Google Scholar]
  • 52.Farrer M, Kachergus J, Forno L, Lincoln S, Wang DS, Hulihan M, Maraganore D, Gwinn-Hardy K, Wszolek Z, Dickson D, Langston JW. Comparison of kindreds with parkinsonism and alpha-synuclein genomic multiplications. Ann. Neurol. 2004;55:174–179. doi: 10.1002/ana.10846. [DOI] [PubMed] [Google Scholar]
  • 53.Singleton AB, Farrer M, Johnson J, Singleton A, Hague S, Kachergus J, Hulihan M, Peuralinna T, Dutra A, Nussbaum R, Lincoln S, Crawley A, Hanson M, Maraganore D, Adler C, Cookson MR, Muenter M, Baptista M, Miller D, Blancato J, Hardy J, Gwinn-Hardy K. alpha-Synuclein locus triplication causes Parkinson's disease. Science. 2003;302:841. doi: 10.1126/science.1090278. [DOI] [PubMed] [Google Scholar]
  • 54.Conway KA, Harper JD, Lansbury PT. Accelerated in vitro fibril formation by a mutant alpha-synuclein linked to early-onset Parkinson disease. Nat. Med. 1998;4:1318–1320. doi: 10.1038/3311. [DOI] [PubMed] [Google Scholar]
  • 55.Giasson BI, Uryu K, Trojanowski JQ, Lee VMY. Mutant and wild type human alpha-synucleins assemble into elongated filaments with distinct morphologies in vitro. J. Biol. Chem. 1999;274:7619–7622. doi: 10.1074/jbc.274.12.7619. [DOI] [PubMed] [Google Scholar]
  • 56.Hashimoto M, Hsu LJ, Sisk A, Xia Y, Takeda A, Sundsmo M, Masliah E. Human recombinant NACP/alpha-synuclein is aggregated and fibrillated in vitro: relevance for Lewy body disease. Brain Res. 1998;799:301–306. doi: 10.1016/s0006-8993(98)00514-9. [DOI] [PubMed] [Google Scholar]
  • 57.Narhi L, Wood SJ, Steavenson S, Jiang Y, Wu GM, Anafi D, Kaufman SA, Martin F, Sitney K, Denis P, Louis JC, Wypych J, Biere AL, Citron M. Both familial Parkinson's disease mutations accelerate alpha-synuclein aggregation. J. Biol. Chem. 1999;274:9843–9846. doi: 10.1074/jbc.274.14.9843. [DOI] [PubMed] [Google Scholar]
  • 58.Uversky VN, Li J, Fink AL. Evidence for a Partially-Folded Intermediate in α-Synuclein fibril formation. J. Biol. Chem. 2001;276:10737–10744. doi: 10.1074/jbc.M010907200. [DOI] [PubMed] [Google Scholar]
  • 59.Wood SJ, Wypych J, Steavenson S, Louis JC, Citron M, Biere AL. alpha-synuclein fibrillogenesis is nucleation-dependent. Implications for the pathogenesis of Parkinson's disease. J. Biol. Chem. 1999;274:19509–19512. doi: 10.1074/jbc.274.28.19509. [DOI] [PubMed] [Google Scholar]
  • 60.Li J, Uversky VN, Fink AL. Effect of familial Parkinson's disease point mutations A30P and A53T on the structural properties, aggregation, and fibrillation of human alpha-synuclein. Biochemistry. 2001;40:11604–11613. doi: 10.1021/bi010616g. [DOI] [PubMed] [Google Scholar]
  • 61.Conway KA, Harper JD, Lansbury PT. Fibrils formed in vitro from alpha-synuclein and two mutant forms linked to Parkinson's disease are typical amyloid. Biochemistry. 2000;39:2552–2563. doi: 10.1021/bi991447r. [DOI] [PubMed] [Google Scholar]
  • 62.Serpell LC, Berriman J, Jakes R, Goedert M, Crowther RA. Fiber diffraction of synthetic alpha-synuclein filaments shows amyloid- like cross-beta conformation. Proc. Natl. Acad. Sci. U. S. A. 2000;97:4897–4902. doi: 10.1073/pnas.97.9.4897. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Kamiyoshihara T, Kojima M, Ueda K, Tashiro M, Shimotakahara S. Observation of multiple intermediates in alpha-synuclein fibril formation by singular value decomposition analysis. Biochem. Biophys. Res. Commun. 2007;355:398–403. doi: 10.1016/j.bbrc.2007.01.162. [DOI] [PubMed] [Google Scholar]
  • 64.Giasson BI, Murray IVJ, Trojanowski JQ, Lee VMY. A hydrophobic stretch of 12 amino acid residues in the middle of alpha-synuclein is essential for filament assembly. J. Biol. Chem. 2001;276:2380–2386. doi: 10.1074/jbc.M008919200. [DOI] [PubMed] [Google Scholar]
  • 65.Biere AL, Wood SJ, Wypych J, Steavenson S, Jiang Y, Anafi D, Jacobsen FW, Jarosinski MA, Wu GM, Louis JC, Martin F, Narhi LO, Citron M. Parkinson's disease-associated alpha-synuclein is more fibrillogenic than β- and γ-synuclein and cannot cross-seed its homologs. J. Biol. Chem. 2000;275:34574–34579. doi: 10.1074/jbc.M005514200. [DOI] [PubMed] [Google Scholar]
  • 66.Miake H, Mizusawa H, Iwatsubo T, Hasegawa M. Biochemical characterization of the core structure of alpha-synuclein filaments. J. Biol. Chem. 2002;277:19213–19219. doi: 10.1074/jbc.M110551200. [DOI] [PubMed] [Google Scholar]
  • 67.Murray IV, Giasson BI, Quinn SM, Koppaka V, Axelsen PH, Ischiropoulos H, Trojanowski JQ, Lee VM. Role of alpha-synuclein carboxy-terminus on fibril formation in vitro. Biochemistry. 2003;42:8530–8540. doi: 10.1021/bi027363r. [DOI] [PubMed] [Google Scholar]
  • 68.Kessler JC, Rochet JC, Lansbury PT., Jr The N-terminal repeat domain of alpha-synuclein inhibits beta-sheet and amyloid fibril formation. Biochemistry. 2003;42:672–678. doi: 10.1021/bi020429y. [DOI] [PubMed] [Google Scholar]
  • 69.Conway KA, Lee SJ, Rochet JC, Ding TT, Williamson RE, Lansbury PT. 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. U. S. A. 2000;97:571–576. doi: 10.1073/pnas.97.2.571. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Conway KA, Lee SJ, Rochet JC, Ding TT, Harper JD, Williamson RE, Lansbury PT. Accelerated oligomerization by Parkinson's disease linked alpha-synuclein mutants. Ann. N. Y. Acad. Sci. 2000;920:42–45. doi: 10.1111/j.1749-6632.2000.tb06903.x. [DOI] [PubMed] [Google Scholar]
  • 71.Greenbaum EA, Graves CL, Mishizen-Eberz AJ, Lupoli MA, Lynch DR, Englander SW, Axelsen PH, Giasson BI. The E46K mutation in alpha -synuclein increases amyloid fibril formation. J. Biol. Chem. 2005;280:7800–7807. doi: 10.1074/jbc.M411638200. [DOI] [PubMed] [Google Scholar]
  • 72.Choi W, Zibaee S, Jakes R, Serpell LC, Davletov B, Crowther RA, Goedert M. Mutation E46K increases phospholipid binding and assembly into filaments of human alpha-synuclein. FEBS Lett. 2004;576:363–368. doi: 10.1016/j.febslet.2004.09.038. [DOI] [PubMed] [Google Scholar]
  • 73.Bussell R, Jr, Eliezer D. Residual structure and dynamics in Parkinson's disease-associated mutants of alpha-synuclein. J. Biol. Chem. 2001;276:45996–46003. doi: 10.1074/jbc.M106777200. [DOI] [PubMed] [Google Scholar]
  • 74.Waxman EA, Duda JE, Giasson BI. Characterization of antibodies that selectively detect alpha-synuclein in pathological inclusions. Acta Neuropathol. 2008;116:37–46. doi: 10.1007/s00401-008-0375-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Jensen PH, Nielsen MS, Jakes R, Dotti CG, Goedert M. Binding of alpha-synuclein to brain vesicles is abolished by familial Parkinson's disease mutation. J. Biol. Chem. 1998;273:26292–26294. doi: 10.1074/jbc.273.41.26292. [DOI] [PubMed] [Google Scholar]
  • 76.Kahle PJ, Neumann M, Ozmen L, Muller V, Jacobsen H, Schindzielorz A, Okochi M, Leimer U, van der PH, Probst A, Kremmer E, Kretzschmar HA, Haass C. 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. doi: 10.1523/JNEUROSCI.20-17-06365.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Goldberg MS, Lansbury PT. Is there a cause-and-effect relationship between alpha-synuclein fibrillization and Parkinson's disease? Nat. Cell Biol. 2000;2:E115–E119. doi: 10.1038/35017124. [DOI] [PubMed] [Google Scholar]
  • 78.Ding TT, Lee SJ, Rochet JC, Lansbury PT., Jr Annular alpha-synuclein protofibrils are produced when spherical protofibrils are incubated in solution or bound to brain-derived membranes. Biochemistry. 2002;41:10209–10217. doi: 10.1021/bi020139h. [DOI] [PubMed] [Google Scholar]
  • 79.Lashuel HA, Hartley D, Petre BM, Walz T, Lansbury PT. Neurodegenerative disease: Amyloid pores from pathogenic mutations. Nature. 2002;418:291. doi: 10.1038/418291a. [DOI] [PubMed] [Google Scholar]
  • 80.Volles MJ, Lee S-J, Rochet J-C, Shtilerman MD, Ding TT, Kessler JC, Lansbury PT. Vesicle permeabilization by protofibrillar α-synuclein: implications for the pathogenesis and treatment of Parkinson's disease. Biochemistry. 2001;40:7812–7819. doi: 10.1021/bi0102398. [DOI] [PubMed] [Google Scholar]
  • 81.Zakharov SD, Hulleman JD, Dutseva EA, Antonenko YN, Rochet JC, Cramer WA. Helical alpha-synuclein forms highly conductive ion channels. Biochemistry. 2007;46:14369–14379. doi: 10.1021/bi701275p. [DOI] [PubMed] [Google Scholar]
  • 82.Tsigelny IF, Bar-On P, Sharikov Y, Crews L, Hashimoto M, Miller MA, Keller SH, Platoshyn O, Yuan JX, Masliah E. Dynamics of alpha-synuclein aggregation and inhibition of pore-like oligomer development by beta-synuclein. FEBS J. 2007;274:1862–1877. doi: 10.1111/j.1742-4658.2007.05733.x. [DOI] [PubMed] [Google Scholar]
  • 83.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. Single particle characterization of iron-induced pore-forming α-synuclein oligomers. J. Biol. Chem. 2008;283:10992–11003. doi: 10.1074/jbc.M709634200. [DOI] [PubMed] [Google Scholar]
  • 84.Kayed R, Sokolov Y, Edmonds B, McIntire TM, Milton SC, Hall JE, Glabe CG. Permeabilization of lipid bilayers is a common conformation-dependent activity of soluble amyloid oligomers in protein misfolding diseases. J. Biol. Chem. 2004;279:46363–46366. doi: 10.1074/jbc.C400260200. [DOI] [PubMed] [Google Scholar]
  • 85.Kanda S, Bishop JF, Eglitis MA, Yang Y, Mouradian MM. Enhanced vulnerability to oxidative stress by alpha-synuclein mutations and C-terminal truncation. Neuroscience. 2000;97:279–284. doi: 10.1016/s0306-4522(00)00077-4. [DOI] [PubMed] [Google Scholar]
  • 86.Ko L, Mehta ND, Farrer M, Easson C, Hussey J, Yen S, Hardy J, Yen SH. Sensitization of neuronal cells to oxidative stress with mutated human alpha-synuclein. J. Neurochem. 2000;75:2546–2554. doi: 10.1046/j.1471-4159.2000.0752546.x. [DOI] [PubMed] [Google Scholar]
  • 87.Lehmensiek V, Tan EM, Schwarz J, Storch A. Expression of mutant alpha-synucleins enhances dopamine transporter-mediated MPP+ toxicity in vitro. Neuroreport. 2002;13:1279–1283. doi: 10.1097/00001756-200207190-00013. [DOI] [PubMed] [Google Scholar]
  • 88.Ostrerova-Golts N, Petrucelli L, Hardy J, Lee JM, Farer M, Wolozin B. The A53T alpha-synuclein mutation increases iron-dependent aggregation and toxicity. J. Neurosci. 2000;20:6048–6054. doi: 10.1523/JNEUROSCI.20-16-06048.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Ostrerova N, Petrucelli L, Farrer M, Mehta N, Choi P, Hardy J, Wolozin B. alpha-Synuclein shares physical and functional homology with 14-3-3 proteins. J. Neurosci. 1999;19:5782–5791. doi: 10.1523/JNEUROSCI.19-14-05782.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Stefanis L, Kholodilov N, Rideout HJ, Burke RE, Greene LA. Synuclein-1 is selectively up-regulated in response to nerve growth factor treatment in PC12 cells. J. Neurochem. 2001;76:1165–1176. doi: 10.1046/j.1471-4159.2001.00114.x. [DOI] [PubMed] [Google Scholar]
  • 91.Tabrizi SJ, Orth M, Wilkinson JM, Taanman JW, Warner TT, Cooper JM, Schapira AH. Expression of mutant alpha-synuclein causes increased susceptibility to dopamine toxicity. Hum. Mol. Genet. 2000;9:2683–2689. doi: 10.1093/hmg/9.18.2683. [DOI] [PubMed] [Google Scholar]
  • 92.Tanaka Y, Engelender S, Igarashi S, Rao RK, Wanner T, Tanzi RE, Sawa A, Dawson L, Dawson TM, Ross CA. Inducible expression of mutant alpha-synuclein decreases proteasome activity and increases sensitivity to mitochondria-dependent apoptosis. Hum. Mol. Genet. 2001;10:919–926. doi: 10.1093/hmg/10.9.919. [DOI] [PubMed] [Google Scholar]
  • 93.Zhou W, Hurlbert MS, Schaack J, Prasad KN, Freed CR. Overexpression of human alpha-synuclein causes dopamine neuron death in rat primary culture and immortalized mesencephalon-derived cells. Brain Res. 2000;866:33–43. doi: 10.1016/s0006-8993(00)02215-0. [DOI] [PubMed] [Google Scholar]
  • 94.Zhou ZD, Yap BP, Gung AY, Leong SM, Ang ST, Lim TM. Dopamine-related and caspase-independent apoptosis in dopaminergic neurons induced by overexpression of human wild type or mutant alpha-synuclein. Exp. Cell Res. 2006;312:156–170. doi: 10.1016/j.yexcr.2005.10.012. [DOI] [PubMed] [Google Scholar]
  • 95.Zhou W, Schaack J, Zawada WM, Freed CR. Overexpression of human alpha-synuclein causes dopamine neuron death in primary human mesencephalic culture. Brain Res. 2002;926:42–50. doi: 10.1016/s0006-8993(01)03292-9. [DOI] [PubMed] [Google Scholar]
  • 96.Smith WW, Jiang H, Pei Z, Tanaka Y, Morita H, Sawa A, Dawson VL, Dawson TM, Ross CA. Endoplasmic reticulum stress and mitochondrial cell death pathways mediate A53T mutant alpha-synuclein-induced toxicity. Hum. Mol. Genet. 2005;14:3801–3811. doi: 10.1093/hmg/ddi396. [DOI] [PubMed] [Google Scholar]
  • 97.Gosavi N, Lee HJ, Lee JS, Patel S, Lee SJ. Golgi fragmentation occurs in the cells with prefibrillar alpha -synuclein aggregates and precedes the formation of fibrillar inclusion. J. Biol. Chem. 2002;277:48984–48992. doi: 10.1074/jbc.M208194200. [DOI] [PubMed] [Google Scholar]
  • 98.Lee FJ, Liu F, Pristupa ZB, Niznik HB. Direct binding and functional coupling of alpha-synuclein to the dopamine transporters accelerate dopamine-induced apoptosis. FASEB J. 2001;15:916–926. doi: 10.1096/fj.00-0334com. [DOI] [PubMed] [Google Scholar]
  • 99.Lee M, Hyun DH, Halliwell B, Jenner P. Effect of the overexpression of wild-type or mutant alpha-synuclein on cell susceptibility to insult. J. Neurochem. 2001;76:998–1009. doi: 10.1046/j.1471-4159.2001.00149.x. [DOI] [PubMed] [Google Scholar]
  • 100.Petrucelli L, O'Farrell C, Lockhart PJ, Baptista M, Kehoe K, Vink L, Choi P, Wolozin B, Farrer M, Hardy J, Cookson MR. Parkin protects against the toxicity associated with mutant alpha- synuclein: proteasome dysfunction selectively affects catecholaminergic neurons. Neuron. 2002;36:1007–1019. doi: 10.1016/s0896-6273(02)01125-x. [DOI] [PubMed] [Google Scholar]
  • 101.Saha AR, Ninkina NN, Hanger DP, Anderton BH, Davies AM, Buchman VL. Induction of neuronal death by alpha-synuclein. Eur. J. Neurosci. 2000;12:3073–3077. doi: 10.1046/j.1460-9568.2000.00210.x. [DOI] [PubMed] [Google Scholar]
  • 102.Xu J, Kao SY, Lee FJ, Song W, Jin LW, Yankner BA. Dopamine-dependent neurotoxicity of alpha-synuclein: a mechanism for selective neurodegeneration in Parkinson disease. Nat. Med. 2002;8:600–606. doi: 10.1038/nm0602-600. [DOI] [PubMed] [Google Scholar]
  • 103.Jensen PJ, Alter BJ, O'Malley KL. Alpha-synuclein protects naive but not dbcAMP-treated dopaminergic cell types from 1-methyl-4-phenylpyridinium toxicity. J Neurochem. 2003;86:196–209. doi: 10.1046/j.1471-4159.2003.01835.x. [DOI] [PubMed] [Google Scholar]
  • 104.Martin FL, Williamson SJ, Paleologou KE, Hewitt R, El-Agnaf OM, Allsop D. Fe(II)-induced DNA damage in alpha-synuclein-transfected human dopaminergic BE(2)-M17 neuroblastoma cells: detection by the Comet assay. J. Neurochem. 2003;87:620–630. doi: 10.1046/j.1471-4159.2003.02013.x. [DOI] [PubMed] [Google Scholar]
  • 105.da Costa CA, Ancolio K, Checler F. Wild-type but not Parkinson's disease-related ala-53 --> Thr mutant alpha -synuclein protects neuronal cells from apoptotic stimuli. J. Biol. Chem. 2000;275:24065–24069. doi: 10.1074/jbc.M002413200. [DOI] [PubMed] [Google Scholar]
  • 106.Wersinger C, Sidhu A. Differential cytotoxicity of dopamine and H2O2 in a human neuroblastoma divided cell line transfected with [alpha]-synuclein and its familial Parkinson's disease-linked mutants. Neurosci. Lett. 2003;342:124–128. doi: 10.1016/s0304-3940(03)00212-x. [DOI] [PubMed] [Google Scholar]
  • 107.Hashimoto M, Hsu LJ, Rockenstein E, Takenouchi T, Mallory M, Masliah E. alpha-Synuclein protects against oxidative stress via inactivation of the c-Jun N-terminal kinase stress-signaling pathway in neuronal cells. J. Biol. Chem. 2002;277:11465–11472. doi: 10.1074/jbc.M111428200. [DOI] [PubMed] [Google Scholar]
  • 108.Alves da Costa C, Paitel E, Vincent B, Checler F. alpha -Synuclein Lowers p53-dependent Apoptotic Response of Neuronal Cells. J. Biol. Chem. 2002;277:50980–50984. doi: 10.1074/jbc.M207825200. [DOI] [PubMed] [Google Scholar]
  • 109.Colapinto M, Mila S, Giraudo S, Stefanazzi P, Molteni M, Rossetti C, Bergamasco B, Lopiano L, Fasano M. alpha-Synuclein protects SH-SY5Y cells from dopamine toxicity. Biochem. Biophys. Res. Commun. 2006;349:1294–1300. doi: 10.1016/j.bbrc.2006.08.163. [DOI] [PubMed] [Google Scholar]
  • 110.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. Alpha-synuclein regulates neuronal survival via Bcl-2 family expression and PI3/Akt kinase pathway. FASEB J. 2002;16:1826–1828. doi: 10.1096/fj.02-0041fje. [DOI] [PubMed] [Google Scholar]
  • 111.Kayed R, Head E, Thompson JL, McIntire TM, Milton SC, Cotman CW, Glabe CG. Common structure of soluble amyloid oligomers implies common mechanism of pathogenesis. Science. 2003;300:486–489. doi: 10.1126/science.1079469. [DOI] [PubMed] [Google Scholar]
  • 112.Liu CW, Giasson BI, Lewis KA, Lee VM, DeMartino GN, Thomas PJ. A Precipitating Role for Truncated α-Synuclein and the Proteasome in {alpha}-Synuclein Aggregation: implications for pathogenesis of Parkinson disease. J. Biol. Chem. 2005;280:22670–22678. doi: 10.1074/jbc.M501508200. [DOI] [PubMed] [Google Scholar]
  • 113.Klegeris A, Giasson BI, Zhang H, Maguire J, Pelech S, McGeer PL. Alpha-synuclein and its disease-causing mutants induce ICAM-1 and IL-6 in human astrocytes and astrocytoma cells. FASEB J. 2006;20:2000–2008. doi: 10.1096/fj.06-6183com. [DOI] [PubMed] [Google Scholar]
  • 114.Tetzlaff JE, Putcha P, Outeiro TF, Ivanov A, Berezovska O, Hyman BT, McLean PJ. CHIP targets toxic α-synuclein oligomers for degradation. J Biol. Chem. 2008;283:17962–17968. doi: 10.1074/jbc.M802283200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Mazzulli JR, Mishizen AJ, Giasson BI, Lynch DR, Thomas SA, Nakashima A, Nagatsu T, Ota A, Ischiropoulos H. Cytosolic catechols inhibit alpha-synuclein aggregation and facilitate the formation of intracellular soluble oligomeric intermediates. J. Neurosci. 2006;26:10068–10078. doi: 10.1523/JNEUROSCI.0896-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Soper JH, Roy S, Stieber A, Lee E, Wilson RB, Trojanowski JQ, Burd CG, Lee VMY. α-Synuclein-induced Aggregation of Cytoplasmic Vesicles in Saccharomyces cerevisiae. Mol. Biol. Cell. 2008;19:1093–1103. doi: 10.1091/mbc.E07-08-0827. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Outeiro TF, Lindquist S. Yeast cells provide insight into alpha-synuclein biology and pathobiology. Science. 2003;302:1772–1775. doi: 10.1126/science.1090439. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Cooper AA, Gitler AD, Cashikar A, Haynes CM, Hill KJ, Bhullar B, Liu K, Xu K, Strathearn KE, Liu F, Cao S, Caldwell KA, Caldwell GA, Marsischky G, Kolodner RD, Labaer J, Rochet JC, Bonini NM, Lindquist S. Alpha-synuclein blocks ER-Golgi traffic and Rab1 rescues neuron loss in Parkinson's models. Science. 313;2006:324–328. doi: 10.1126/science.1129462. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Flower TR, Chesnokova LS, Froelich CA, Dixon C, Witt SN. Heat Shock Prevents Alpha-synuclein-induced apoptosis in a yeast model of Parkinson's disease. J. Mol. Biol. 2005;351:1081–1100. doi: 10.1016/j.jmb.2005.06.060. [DOI] [PubMed] [Google Scholar]
  • 120.Gitler AD, Bevis BJ, Shorter J, Strathearn KE, Hamamichi S, Su LJ, Caldwell KA, Caldwell GA, Rochet JC, McCaffery JM, Barlowe C, Lindquist S. The Parkinson's disease protein α-synuclein disrupts cellular Rab homeostasis. Proc. Natl. Acad. Sci. U. S. A. 2008;105:145–150. doi: 10.1073/pnas.0710685105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Feany MB, Bender WW. A Drosophila model of Parkinson's disease. Nature. 2000;404:394–398. doi: 10.1038/35006074. [DOI] [PubMed] [Google Scholar]
  • 122.Auluck PK, Chan E, Trojanoswki JQ, Lee VMY, Bonini NM. Chaperon suppression of α-synuclein toxicity in a Drosophila model of Parkinson's disease. Science. 2002;295:865–868. doi: 10.1126/science.1067389. [DOI] [PubMed] [Google Scholar]
  • 123.Pesah Y, Burgess H, Middlebrooks B, Ronningen K, Prosser J, Tirunagaru V, Zysk J, Mardon G. Whole-mount analysis reveals normal numbers of dopaminergic neurons following misexpression of alpha-Synuclein in Drosophila. Genesis. 2005;41:154–159. doi: 10.1002/gene.20106. [DOI] [PubMed] [Google Scholar]
  • 124.Lakso M, Vartianinen S, Moilanen AM, Sirvio J, Thomas JH, Nass R, Blakely RD, Wong G. Dopaminergic neuronal loss and motor deficits in Caenorhabdtis elegans overexpressing human alpha-synuclein. J. Neurochem. 2003;86:165–172. doi: 10.1046/j.1471-4159.2003.01809.x. [DOI] [PubMed] [Google Scholar]
  • 125.Kuwahara T, Koyama A, Gengyo-Ando K, Masuda M, Kowa H, Tsunoda M, Mitani S, Iwatsubo T. Familial Parkinson mutant alph-synuclein causes dopamine neuron dysfunction in transgenic Caenorhabditis elegans. J. Biol. Chem. 2006;281:334–340. doi: 10.1074/jbc.M504860200. [DOI] [PubMed] [Google Scholar]
  • 126.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 alpha-synuclein mice: implications for neurodegenerative disorders. Science. 2000;287:1265–1269. doi: 10.1126/science.287.5456.1265. [DOI] [PubMed] [Google Scholar]
  • 127.Hashimoto M, Rockenstein E, Mante M, Mallory M, Masliah E. β-synuclein inhibits α-synuclein aggregation: a possible role as an anti-parkinsonian factor. Neuron. 2001;32:213–223. doi: 10.1016/s0896-6273(01)00462-7. [DOI] [PubMed] [Google Scholar]
  • 128.Masliah E, Rockenstein E, Veinbergs I, Sagara Y, Mallory M, Hashimoto M, Mucke L. beta-Amyloid peptides enhance alpha-synuclein accumulation and neuronal deficits in a transgenic mouse model linking Alzheimer's disease and Parkinson's disease. Proc. Natl. Acad. Sci. U. S. A. 2001;98:12245–12250. doi: 10.1073/pnas.211412398. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Klucken J, Shin Y, Masliah E, Hyman BT, McLean PJ. Hsp70 reduces alpha-synuclein aggregation and toxicity. J. Biol. Chem. 2004;279:25497–25502. doi: 10.1074/jbc.M400255200. [DOI] [PubMed] [Google Scholar]
  • 130.van der Putten H, Wiederhold KH, Probst A, Barbieri S, Mistl C, Danner S, Kauffmann S, Hofele K, Spooren WPJM, Ruegg MA, Lin S, Caroni P, Sommer B, Tolnay M, Bilbe G. Neuropathology in mice expressing human alpha-synuclein. J. Neurosci. 2000;20:6021–6029. doi: 10.1523/JNEUROSCI.20-16-06021.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Kahle PJ, Neumann M, Ozmen L, Muller V, Odoy S, Okamoto N, Jacobsen H, Iwatsubo T, Trojanoswki JQ, Takahashi H, Wakabayashi K, Bogdanovic N, Riedered P, Kretzschmar HA, Haass C. Selective insolubility of α-synuclein in human Lewy body diseases is recapitalated in a transgenic mouse model. Am. J. Pathol. 2001;159:2215–2225. doi: 10.1016/s0002-9440(10)63072-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.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. Misfolded proteinase K-resistant hyperphosphorylated alpha-synuclein in aged transgenic mice with locomotor deterioration and in human alpha-synucleinopathies. J. Clin. Invest. 2002;110:1429–1439. doi: 10.1172/JCI15777. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Freichel C, Neumann M, Ballard T, Mnller V, Woolley M, Ozmen L, Borroni E, Kretzschmar HA, Haass C, Spooren W, Kahle PJ. Age-dependent cognitive decline and amygdala pathology in α-synuclein transgenic mice. Neurobiol. Aging. 2007;28:1421–1435. doi: 10.1016/j.neurobiolaging.2006.06.013. [DOI] [PubMed] [Google Scholar]
  • 134.Giasson BI, Duda JE, Quinn SM, Zhang B, Trojanoswki JQ, Lee VMY. Neuronal α-synucleinopathy with severe movement disorder in mice expressing A53T human α-synuclein. Neuron. 2002;34:521–533. doi: 10.1016/s0896-6273(02)00682-7. [DOI] [PubMed] [Google Scholar]
  • 135.Lee MK, Stirling W, Xu Y, Xu X, Qui D, Mandir AS, Dawson TM, Copeland NG, Jenkins NA, Price DL. Human alpha-synuclein-harboring familial Parkinson's disease-linked Ala-53 --> Thr mutation causes neurodegenerative disease with alpha-synuclein aggregation in transgenic mice. Proc. Natl. Acad. Sci. U. S. A. 2002;99:8968–8973. doi: 10.1073/pnas.132197599. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Martin LJ, Pan Y, Price AC, Sterling W, Copeland NG, Jenkins NA, Price DL, Lee MK. Parkinson's disease α-synuclein transgenic mice develop neuronal mitochondrial degeneration and cell death. J. Neurosci. 2006;26:41–50. doi: 10.1523/JNEUROSCI.4308-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Gallardo G, Schluter OM, Sudhof TC. A molecular pathway of neurodegeneration linking alpha-synuclein to ApoE and Abeta peptides. Nat. Neurosci. 2008;11:301–308. doi: 10.1038/nn2058. [DOI] [PubMed] [Google Scholar]
  • 138.Ihara M, Tomimoto H, Kitayama H, Morioka Y, Akiguchi I, Shibasaki H, Noda M, Kinoshita M. Association of the cytoskeletal GTP-binding protein Sept4/H5 with cytoplasmic inclusions found in Parkinson's disease and other synucleinopathies. J. Biol. Chem. 2003;278:24095–24102. doi: 10.1074/jbc.M301352200. [DOI] [PubMed] [Google Scholar]
  • 139.Ihara M, Yamasaki N, Hagiwara A, Tanigaki A, Kitano A, Hikawa R, Tomimoto H, Noda M, Takanashi M, Mori H, Hattari N, Miyakawa T, Kinoshita M. Sept4, a component of presynaptic scaffold and Lewy bodies, is required for the suppression of alpha-synuclein neurotoxicity. Neuron. 2007;53:519–533. doi: 10.1016/j.neuron.2007.01.019. [DOI] [PubMed] [Google Scholar]
  • 140.Matsuoka Y, Vila M, Lincoln S, McCormack A, Picciano M, LaFrancois J, Yu X, Dickson D, Langston WJ, McGowan E, Farrer M, Hardy J, Duff K, Przedborski S, Di Monte DA. Lack of nigral pathology in transgenic mice expressing human alpha-synuclein driven by the tyrosine hydroxylase promoter. Neurobiol. Dis. 2001;8:535–539. doi: 10.1006/nbdi.2001.0392. [DOI] [PubMed] [Google Scholar]
  • 141.Baba M, Nakajo S, Tu PH, Tomita T, Nakaya K, Lee VMY, Trojanowski JQ, Iwatsubo T. Aggregation of alpha-synuclein in Lewy bodies of sporadic Parkinson's disease and dementia with Lewy bodies. Am. J. Pathol. 1998;152:879–884. [PMC free article] [PubMed] [Google Scholar]
  • 142.Campbell BC, McLean CA, Culvenor JG, Gai WP, Blumbergs PC, Jakala P, Beyreuther K, Masters CL, Li QX. The solubility of alpha-synuclein in multiple system atrophy differs from that of dementia with Lewy bodies and Parkinson's disease. J. Neurochem. 2001;76:87–96. doi: 10.1046/j.1471-4159.2001.00021.x. [DOI] [PubMed] [Google Scholar]
  • 143.Li W, West N, Colla E, Pletnikova O, Troncoso JC, Marsh L, Dawson TM, Jakala P, Hartmann T, Price DL. Aggregation promoting C-terminal truncation of α-synuclein is a normal cellular process and is enhanced by the familial Parkinson's disease-linked mutations. Proc. Natl. Acad. Sci. U. S. A. 2005;102:2162–2167. doi: 10.1073/pnas.0406976102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Mishizen-Eberz AJ, Guttmann RP, Giasson BI, Day GA, III, Hodara R, Ischiropoulos H, Lee VM, Trojanowski JQ, Lynch DR. Distinct cleavage patterns of normal and pathologic forms of alpha-synuclein by calpain I in vitro. J. Neurochem. 2003;86:836–847. doi: 10.1046/j.1471-4159.2003.01878.x. [DOI] [PubMed] [Google Scholar]
  • 145.Dufty BM, Warner LR, Hou ST, Jiang SX, Gomez-Isla T, Leenhouts KM, Oxford JT, Feany MB, Masliah E, Rohn TT. Calpain-cleavage of alpha-synuclein: connecting proteolytic processing to disease-linked aggregation. Am. J Pathol. 2007;170:1725–1738. doi: 10.2353/ajpath.2007.061232. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Tofaris GK, Garcia Reitbock P, Humby T, Lambourne SL, O'Connell M, Ghetti B, Gossage H, Emson PC, Wilkinson LS, Goedert M, Grazia Spillantini M. Pathological changes in dopaminergic nerve cells of the substantia nigra and olfactory bulb in mice transgenic for truncated human α-synuclein(1–120): Implications for Lewy body disorders. J. Neurosci. 2006;26:3942–3950. doi: 10.1523/JNEUROSCI.4965-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Wakamatsu M, Ishii A, Iwata S, Sakagami J, Ukai Y, Ono M, Kanbe D, Muramatsu Si, Kobayashi K, Iwatsubo T, Yoshimoto M. Selective loss of nigral dopamine neurons induced by overexpression of truncated human α-synuclein in mice. Neurobiol. Aging. 2008;29:574–585. doi: 10.1016/j.neurobiolaging.2006.11.017. [DOI] [PubMed] [Google Scholar]
  • 148.Kahle PJ, Neumann M, Ozmen L, Muller V, Jacobsen H, Spooren W, Fuss B, Mallon B, Macklin WB, Fujiwara H, Hasegawa M, Iwatsubo T, Kretzschmar HA, Haass C. Hyperphosphorylation and insolubility of alpha-synuclein in transgenic mouse oligodendrocytes. EMBO Rep. 2002;3:583–588. doi: 10.1093/embo-reports/kvf109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Stefanova N, Reindl M, Neumann M, Haass C, Poewe W, Kahle PJ, Wenning GK. Oxidative stress in transgenic mice with oligodendroglial α-synuclein overexpression replicates the characteristic neuropathology of multiple system atrophy. Am. J. Pathol. 2005;166:869–876. doi: 10.1016/s0002-9440(10)62307-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Shults CW, Rockenstein E, Crews L, Adame A, Mante M, Larrea G, Hashimoto M, Song D, Iwatsubo T, Tsuboi K, Masliah E. Neurological and Neurodegenerative Alterations in a transgenic mouse model expressing human α-synuclein under oligodendrocyte promoter: Implications for multiple system atrophy. J. Neurosci. 2005;25:10689–10699. doi: 10.1523/JNEUROSCI.3527-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Yazawa I, Giasson BI, Sasaki R, Zhang B, Joyce S, Uryu K, Trojanowski JQ, Lee VMY. Mouse model of multiple system atrophy α-synuclein expression in oligodendrocytes causes glial and neuronal degeneration. Neuron. 2005;45:847–859. doi: 10.1016/j.neuron.2005.01.032. [DOI] [PubMed] [Google Scholar]
  • 152.Nuber S, Petrasch-Perwez E, Winner B, Winkler J, vonHorsten S, Schmidt T, Boy J, Kuhn M, Nguyen HP, Teismann P, Schulz JB, Neumann M, Pichler BJ, Reischl G, Holzmann C, Schmitt I, Bornemann A, Kuhn W, Zimmermann F, Servadio A, Riess O. Neurodegeneration and motor dysfunction in a conditional model of Parkinson's disease. J. Neurosci. 2008;25:2471–2484. doi: 10.1523/JNEUROSCI.3040-07.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Kirik D, Rosenblad C, Burger C, Lundberg C, Johansen TE, Muzyczka N, Mandel RJ, Bjorklund A. Parkinson-like neurodegeneration induced by targeted overexpression of alpha -synuclein in the nigrostriatal System. J. Neurosci. 2002;22:2780–2791. doi: 10.1523/JNEUROSCI.22-07-02780.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Kirik D, Annett LE, Burger C, Muzyczka N, Mandel RJ, Bjorklund A. Nigrostriatal α-synucleinopathy induced by viral vector-mediated overexpression of human α-synuclein: A new primate model of Parkinson's disease. Proc. Natl. Acad. Sci. U. S. A. 2003;100:2884–2889. doi: 10.1073/pnas.0536383100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Klein RL, King MA, Hamby ME, Meyer EM. Dopaminergic Cell Loss Induced by Human A30P alpha-Synuclein Gene Transfer to the Rat Substantia Nigra. Hum. Gene Ther. 2002;13:605–612. doi: 10.1089/10430340252837206. [DOI] [PubMed] [Google Scholar]
  • 156.Lo Bianco C, Ridet JL, Schneider BL, Deglon N, Aebischer P. α-Synucleinopathy and selective dopaminergic neuron loss in a rat lentiviral-based model of Parkinson's disease. Proc. Natl. Acad. Sci. U. S. A. 2002;99:10813–10818. doi: 10.1073/pnas.152339799. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Ross OA, Braithwaite AT, Skipper LM, Kachergus J, Hulihan MM, Middleton FA, Nishioka K, Fuchs J, Gasser T, Maraganore DM, Adler CH, Larvor L, Chartier-Harlin MC, Nilsson C, Langston JW, Gwinn K, Hattori N, Farrer MJ. Genomic investigation of alpha-synuclein multiplication and parkinsonism. Ann. Neurol. 2008;63:743–750. doi: 10.1002/ana.21380. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Dachsel JC, Lincoln SJ, Gonzalez J, Ross OA, Dickson DW, Farrer MJ. The ups and downs of alpha-synuclein mRNA expression. Mov. Disord. 2007;22:293–295. doi: 10.1002/mds.21223. [DOI] [PubMed] [Google Scholar]
  • 159.Haroutunian V, Serby M, Purohit DP, Perl DP, Marin D, Lantz M, Mohs RC, Davis KL. Contribution of Lewy body inclusions to dementia in patients with and without Alzheimer disease neuropathological conditions. Arch. Neurol. 2000;57:1145–1150. doi: 10.1001/archneur.57.8.1145. [DOI] [PubMed] [Google Scholar]
  • 160.Hurtig HI, Trojanowski JQ, Galvin J, Ewbank D, Schmidt ML, Lee VM, Clark CM, Glosser G, Stern MB, Gollomp SM, Arnold SE. Alpha-synuclein cortical Lewy bodies correlate with dementia in Parkinson's disease. Neurology. 2000;54:1916–1921. doi: 10.1212/wnl.54.10.1916. [DOI] [PubMed] [Google Scholar]
  • 161.Lennox G, Lowe J, Landon M, Byrne EJ, Mayer RJ, Godwin-Austen RB. Diffuse Lewy body disease: correlative neuropathology using anti- ubiquitin immunocytochemistry. J. Neurol. Neurosurg. Psychiatry. 1989;52:1236–1247. doi: 10.1136/jnnp.52.11.1236. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Samuel W, Galasko D, Masliah E, Hansen LA. Neocortical lewy body counts correlate with dementia in the Lewy body variant of Alzheimer's disease. J. Neuropathol. Exp. Neurol. 1996;55:44–52. doi: 10.1097/00005072-199601000-00005. [DOI] [PubMed] [Google Scholar]
  • 163.Duda JE, Giasson BI, Mabon ME, Lee VMY, Trojanoswki JQ. Novel antibodies to synuclein show abundant striatal pathology in Lewy body diseases. A nn. Neurol. 2002;52:205–210. doi: 10.1002/ana.10279. [DOI] [PubMed] [Google Scholar]
  • 164.Kramer ML, Schulz-Schaeffer WJ. Presynaptic alpha-synuclein aggregates, not Lewy bodies, cause neurodegeneration in dementia with Lewy bodies. J. Neurosci. 2007;27:1405–1410. doi: 10.1523/JNEUROSCI.4564-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165.Duda JE, Giasson BI, Lee VMY, Trojanowski JQ. Is the initial insult in Parkinson's disease and dementia with Lewy bodies a neuritic dystrophy? Ann. N. Y. Acad. Sci. 2003;991:295–297. [Google Scholar]
  • 166.Lindersson E, Beedholm R, Hojrup P, Moos T, Gai W, Hendil KB, Jensen PH. Proteasomal inhibition by alpha-synuclein filaments and oligomers. J. Biol. Chem. 2004;279:12924–12934. doi: 10.1074/jbc.M306390200. [DOI] [PubMed] [Google Scholar]

RESOURCES